01.08.2013 Views

A Century of Ramjet Propulsion Technology Evolution - Faculty of ...

A Century of Ramjet Propulsion Technology Evolution - Faculty of ...

A Century of Ramjet Propulsion Technology Evolution - Faculty of ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

JOURNAL OF PROPULSION AND POWER<br />

Vol. 20, No. 1, January–February 2004<br />

A <strong>Century</strong> <strong>of</strong> <strong>Ramjet</strong> <strong>Propulsion</strong> <strong>Technology</strong> <strong>Evolution</strong><br />

Ronald S. Fry<br />

Johns Hopkins University, Columbia, Maryland 21044<br />

A general review is presented <strong>of</strong> the worldwide evolution <strong>of</strong> ramjet propulsion since the Wright brothers first<br />

turned man’s imagination to fly into a practical reality. A perspective <strong>of</strong> the technological developments from<br />

subsonic to hypersonic flight speeds is provided to allow an appreciation for the advances made internationally<br />

from the early 1900s to current times. <strong>Ramjet</strong>, scramjet, and mixed-cycle engine types, and their operation and<br />

rationale for use are considered. The development history and principal contributing development programs are<br />

reviewed. Major airbreathing technologies that had significant impact on the maturation <strong>of</strong> ramjet propulsion<br />

and enabled engine designs to mature to their current state are identified. The general state <strong>of</strong> flight-demonstrated<br />

technology is summarized and compared with the technology base <strong>of</strong> 1980. The current status <strong>of</strong> ramjet/scramjet<br />

technology is identified. <strong>Ramjet</strong> and scramjet propulsion technology has matured dramatically over the years<br />

in support <strong>of</strong> both military and space access applications, yet many opportunities remain to challenge future<br />

generations <strong>of</strong> explorers.<br />

Introduction<br />

HISTORY and technology reviews are written and read for many<br />

practical reasons, 1 including their usefulness to managers and<br />

engineers engaged in the advanced technology developments. Although<br />

history does not repeat itself, similar situations <strong>of</strong>ten have<br />

similar results, and thoughtful study <strong>of</strong> past technology developments<br />

can help us to recognize and avoid pitfalls to make desired<br />

outcomes more likely. In any event, study <strong>of</strong> history lets us see current<br />

problems more clearly. Because <strong>of</strong> the cyclic nature <strong>of</strong> research<br />

and developments, it is remarkable how many times something had<br />

to be rediscovered. This is a real and costly problem, and without a<br />

concerted effort to avoid it, it is apt to get worse in the future. Solutions<br />

have been suggested, which include staying current in our<br />

respective fields, capitalizing on making the knowledge explosion<br />

more tractable, and making new technology available to industry.<br />

Einstein ably characterized advances in technology as “If I have<br />

seen farther than others, it is because I stand on the shoulders <strong>of</strong><br />

giants.” Thus, this historical paper draws on the many outstanding<br />

chronicles <strong>of</strong> selected elements <strong>of</strong> airbreathing propulsion development,<br />

assembled and presented for your appreciation <strong>of</strong> how far the<br />

history <strong>of</strong> ramjet technology has come in the last 100 years. <strong>Ramjet</strong><br />

technology has evolved from the early simple subsonic “flying<br />

stovepipe” to its role in the complex combined or mixed cycle concepts<br />

embedded within military and space access vehicle designs<br />

<strong>of</strong> today. This evolution spans far more than just years; it is also a<br />

vast revolution in development technologies. A brief review <strong>of</strong> this<br />

evolution, is provided; a rigorous presentation is beyond the scope<br />

<strong>of</strong> this manuscript.<br />

The history <strong>of</strong> man’s desire to fly has evolved from the imagination<br />

<strong>of</strong> Greek mythologists through the thoughts, designs, data,<br />

and experiences <strong>of</strong> many notable individuals, all whom contributed<br />

in various ways to the Wright brothers’ first demonstration <strong>of</strong> level<br />

flight under power in 1903. It would seem this was the catalyst that<br />

initiated the evolution <strong>of</strong> ramjet technology because the first known<br />

reference to ramjet propulsion dates from 1913.<br />

Review <strong>of</strong> the worldwide evolution <strong>of</strong> ramjet technology begins<br />

with ramjet engine types, operation, and motivation. <strong>Ramjet</strong> development<br />

history is reviewed. Principal international development<br />

programs are reviewed. Discussions are concluded with a review <strong>of</strong><br />

the maturation <strong>of</strong> ramjet design technology, and the general state <strong>of</strong><br />

ramjet/scramjet/mixed cycle technology.<br />

Key enabling technologies, components, or events that had significant<br />

impact on the maturation <strong>of</strong> ramjet and scramjet propulsion<br />

and engine designs are summarized in Table 1. These significant<br />

elements are discussed in further detail throughout the paper and<br />

include the air induction system, vehicle aerodynamics, combustor<br />

design and materials, fuels, injection and mixing, solid propellant<br />

boosters, ejectable and nonejectable components, thermochemical<br />

and engine performance modeling, and ground-test facilities and<br />

methods.<br />

Ronald S. Fry has made contributions to combustion and propulsion technology for the development <strong>of</strong> advanced<br />

airbreathing military applications for over 30 years. He has made contributions to the development <strong>of</strong> conventional<br />

stand<strong>of</strong>f missiles, cluster weapon systems involving self-forging fragment technology, and military weapon systems<br />

involving fuel–air explosive and insensitive high-explosive technologies while in the U.S. Air Force from 1974 to<br />

1979. He led and contributed to dramatic advances in the U.S. state <strong>of</strong> the art in metallized solid fuel ramjet<br />

technology while with Atlantic Research Corporation from 1979 to 1993. More recently, he has contributed to the<br />

development <strong>of</strong> a U.S. national plan for the advancement <strong>of</strong> hypersonics science and technology. R. S. Fry is currently<br />

a Senior Research Engineer at Johns Hopkins University, Chemical <strong>Propulsion</strong> Information Agency, a Department<br />

<strong>of</strong> Defense Information Analysis Center. He has received awards for outstanding sustained achievement from the<br />

JANNAF Airbreathing <strong>Propulsion</strong> Subcommittee, the National Defense Service Medal for Active Duty Service<br />

during the Vietnam War, and multiple U.S. Air Force commendation medals for outstanding performance. He<br />

has a B.S. in aerospace engineering and two M.S. in engineering in aerospace engineering and naval architecture<br />

and marine engineering from the University <strong>of</strong> Michigan. R. S. Fry performed his graduate work under the<br />

mentorship <strong>of</strong> A. Nicholls, whose early demonstration <strong>of</strong> stabilized detonation waves in a supersonic hydrogen–air<br />

stream contributed to the feasibility <strong>of</strong> supersonic combustion ramjet technology. He is a Senior Member <strong>of</strong> AIAA.<br />

Received 25 August 2003; revision received 20 November 2003; accepted for publication 19 November 2003. Copyright c○ 2003 by Ronald S. Fry. Published<br />

by the American Institute <strong>of</strong> Aeronautics and Astronautics, Inc., with permission. Copies <strong>of</strong> this paper may be made for personal or internal use, on condition<br />

that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 0748-4658/04<br />

$10.00 in correspondence with the CCC.<br />

27


28 FRY<br />

Table 1 Top 10 enabling technologies for ramjet propulsion<br />

1) High speed aerodynamics analysis<br />

CFD code analysis and validation methodologies<br />

(external and internal flow)<br />

Improved design tools and techniques<br />

2) Air induction system technology<br />

Fixed and variable geometry<br />

Subsonic, internally/externally ducted supersonic<br />

and dual-flowpath designs<br />

Mixed cycle flowpath development<br />

Improved design tools/integration with the airframe<br />

Improved materials, especially in the cowl region<br />

3) Combustor technology<br />

Improved design tools and techniques, such as<br />

mapping fuel and heat-transfer distributions<br />

Improved insulators (ablative, nonablative)<br />

Advanced structural materials<br />

Combustion ignition, piloting and flameholding, and mixing<br />

4) <strong>Ramjet</strong>/scramjet fuels<br />

Higher-energy liquid and solid fuels<br />

Low-temperature liquid fuels<br />

Endothermic fuels<br />

5) Fuel management systems<br />

Liquid fuel injection and mixing<br />

Improved injectors; wider range <strong>of</strong> operation,<br />

tailoring <strong>of</strong> atomization, and spray distribution<br />

Solid ramjet and ducted rocket fuel grain design<br />

Solid ducted rocket fuel value design<br />

Variable-geometry injection systems, especially for DR<br />

Improved feed systems, including turbopumps<br />

Improved feedback control systems<br />

6) <strong>Propulsion</strong>/airframe integration, materials, and thermal management<br />

CFD code analysis and validation methodologies<br />

High-temperature metals and alloys<br />

High-temperature structures<br />

Passive and active cooling<br />

Carbon–carbon and ceramic metal matrix composites<br />

7) Solid propellant booster technology<br />

Tandem boosters<br />

Integral rocket–ramjet boosters<br />

Self-boosted ramjet (mixed cycle RBCC, TBCC, etc.)<br />

8) Ejectable and nonejectable component technology<br />

Inlet and port covers<br />

Fixed- and variable-geometry nozzle technology<br />

9) Thermochemical modeling and simulation development<br />

Thermochemical tables<br />

<strong>Ramjet</strong> cycle analysis and performance modeling<br />

10) Ground-test methodologies<br />

Direct-connect<br />

Semifreejet and freejet<br />

Airflow quality improvements<br />

Instrumentation advances<br />

Computational tools and flight-test correlation<br />

<strong>Ramjet</strong> Design and Concept <strong>of</strong> Operation<br />

Description <strong>of</strong> <strong>Ramjet</strong> <strong>Propulsion</strong> Systems<br />

Simple in concept, the ramjet uses fixed components to compress<br />

and accelerate intake air by ram effect. The ramjet has been called<br />

a flying stovepipe, due to the absence <strong>of</strong> rotating parts that characterize<br />

the turbine engine. The ramjet gets its name from the method<br />

<strong>of</strong> air compression because it cannot operate from a standing start<br />

but must first be accelerated to a high speed by another means <strong>of</strong><br />

propulsion. The air enters the inlet and diffuser, which serve the<br />

same purpose as a compressor. Compression depends on velocity<br />

and increases dramatically with vehicle speed. The air delivered to<br />

a combustion chamber is mixed with injected fuel. This mixture is<br />

ignited and burns in the presence and aid <strong>of</strong> a flameholder that stabilizes<br />

the flame. The burning fuel imparts thermal energy to the gas,<br />

which expands to high velocity through the nozzle at speeds greater<br />

than the entering air, which produces forward thrust. Because thrust<br />

strongly depends on compression, the ramjet needs forward velocity<br />

to start the cycle. Typically a booster rocket provides this, either<br />

externally or internally. All modern ramjet missiles use the integral<br />

rocket–ramjet concept, which involves solid propellant in the<br />

Fig. 1 Elements <strong>of</strong> ramjet power cycle and flowpath.<br />

aft combustion or mixing chamber to boost the system to ramjetoperating<br />

conditions. On decay <strong>of</strong> rocket pressure, the nozzle and<br />

associated components are jettisoned and ramjet power begins. Low<br />

ejecta booster design trades involve using the less efficient ramjet<br />

nozzle in a trade<strong>of</strong>f for volume, performance, and operational considerations.<br />

Elements <strong>of</strong> the ramjet power cycle and flowpath from<br />

Avery2 and Thomas3 are shown in Fig. 1. Note that variation in<br />

station nomenclature began early.<br />

A typical Mach number–altitude airbreathing flight corridor is<br />

shown4 in Fig. 2. Design challenges are compounded by flight conditions<br />

that become increasingly severe due to the combination <strong>of</strong><br />

internal duct pressure, skin temperature, and dynamic pressure loading.<br />

These constraints combine to create a narrow corridor <strong>of</strong> possible<br />

conditions suitable for flight based on ram air compression. Relatively<br />

high dynamic pressure q is required, compared to a rocket, to<br />

provide adequate static pressure in the combustor (generally more<br />

than 1<br />

2<br />

atm) for good combustion and to provide sufficient thrust.<br />

As speed increases, there is less need for mechanical compression.<br />

The upper boundary is characterized as a region <strong>of</strong> low combustion<br />

efficiency and narrow fuel/air ratio ranges thereby establishing<br />

a combustion limit. The lower boundary is a region <strong>of</strong> high skin<br />

temperature and pressure loading thereby establishing design and<br />

material limits. The far right, high Mach number edge <strong>of</strong> the envelope<br />

is a region <strong>of</strong> extreme dissociation, where nonequilibrium flow<br />

can influence compression ramp flow, induce large leading-edge<br />

heating rates, and create distortions in the inlet flow, while influencing<br />

fuel injection and mixing, combustion chemistry, nozzle flow,<br />

and, ultimately, performance. A region <strong>of</strong> low compression ratio is<br />

experienced at the very low Mach number region. The current state<br />

<strong>of</strong> the ramjet operational envelope, examined in further detail in the<br />

final section, has seen dramatic expansions from its early history,<br />

through the 1980s, to today.<br />

The basic ramjet engine consists <strong>of</strong> an inlet, diffuser, combustor<br />

and exhaust nozzle (Fig. 1). The inlet collects and compresses air<br />

and conducts it via the diffuser to the combustor at reduced velocity<br />

thereby developing ram pressure and an elevated temperature. The<br />

combustor adds heat and mass to the compressed air by the injection<br />

and burning <strong>of</strong> fuel. Finally, the nozzle converts some <strong>of</strong> the energy<br />

<strong>of</strong> the hot combustion products to kinetic energy to produce thrust.<br />

Because the ramjet depends only on its forward motion to compress<br />

the air, the engine itself has no moving parts and <strong>of</strong>fers higher Mach<br />

number capability than turbojet engines. However, unlike a turbojet<br />

or rocket engine, the ramjet requires an auxiliary boost system to<br />

accelerate it to its supersonic operating regime.<br />

There are numerous reasons for using airbreathing engines instead<br />

<strong>of</strong> rockets: All <strong>of</strong> the oxidizer necessary for combustion <strong>of</strong> the


fuels comes from the atmosphere, the engines produce much higher<br />

engine efficiency over a large portion <strong>of</strong> the flight and a longer<br />

powered range than rockets, there is thrust modulation for efficient<br />

cruise and acceleration, they have the ability to change efficiently<br />

powered fight path and maneuverability, and they are reusable not<br />

just refurbishable. An additional feature for space access is a short<br />

turnaround time, with a potential cost reduction <strong>of</strong> 10–100 times per<br />

pound <strong>of</strong> payload.<br />

Possible applications for scramjets include hypersonic cruise vehicles,<br />

hypervelocity missiles, and airbreathing boosters for space<br />

applications. Hydrogen fuel is desirable for high Mach number applications<br />

due to its high-energy content, fast reactions, and excellent<br />

cooling capabilities. For hypersonic missile applications and airbreathing<br />

systems operating below Mach 6, hydrocarbon fuels are<br />

preferred because <strong>of</strong> volumetric and operational constraints. Making<br />

use <strong>of</strong> the enhanced cooling capabilities <strong>of</strong> endothermic hydrocarbon<br />

fuels can increase the maximum speed for hydrocarbon-fueled<br />

missiles and vehicles to Mach 8. Attractive mission identified for<br />

scramjet-powered vehicles include a Mach 8 cruise missile as a<br />

stand<strong>of</strong>f fast-reaction weapon or long-range cruise missile or boost<br />

propulsion for stand<strong>of</strong>f fast-reaction weapon or airbreathing booster<br />

for space access.<br />

Subsonic Combustion (<strong>Ramjet</strong>) Cycle Behavior<br />

Conceptual schematics <strong>of</strong> subsonic combustion ramjets and hybrid<br />

or combined cycle derivatives are shown in Figs. 3 and 4 following<br />

Waltrup, 5 whose past contributions have been most noteworthy.<br />

Figure 3a shows the traditional can-type ramjet (CRJ), liquid-fueled<br />

ramjet (LFRJ), and gaseous-fueled ramjet (GFRJ) with a tandem<br />

booster attached. A tandem booster is required to provide static and<br />

low-speed thrust, which pure ramjets alone cannot provide. Here,<br />

M0 > M1 > 1, but the air is diffused to a subsonic speed (typically<br />

Mach 0.3–0.4) through a normal shock system before reaching station<br />

4. Fuel is then injected and burned with the air at low subsonic<br />

speeds before reaccelerating through a geometric throat (M5 = 1)<br />

and exit nozzle (M6 > 1). The position <strong>of</strong> the normal shock system<br />

in this and all subsonic combustion ramjets is determined by the<br />

flight speed, air captured, total pressure losses up to the inlet’s terminal<br />

normal shock, amount <strong>of</strong> heat addition, inlet boundary layer<br />

and flow distortion, and exit nozzle throat size.<br />

A more recent alternative to this concept is to use a common<br />

combustion chamber, commonly referred to as an integral rocket<br />

ramjet (IRR), for both the boost and sustain phases <strong>of</strong> flight. This<br />

generally requires a dump-type rather than a can-type combustor,<br />

but the cycle <strong>of</strong> operation <strong>of</strong> the ramjet remains the same. Figure 3b<br />

is a schematic <strong>of</strong> this concept for a liquid-fueled IRR (LFIRR) and<br />

Fig. 3c shows a solid-fueled IRR (SFIRR). In some applications,<br />

SFIRRs are preferred over LFIRRs, GFRJs, or CRJs because <strong>of</strong><br />

the simplicity <strong>of</strong> the fuel supply, but only when the fuel throttling<br />

requirements are minimal, that is, when flight altitude and Mach<br />

FRY 29<br />

Fig. 2 Typical airbreathing flight corridor.<br />

a) Conventional can combustor ramjet (CRJ)<br />

b) Integral rocket/dump combustor ramjet (LFIRR)<br />

c) SFIRR<br />

d) ADR<br />

Fig. 3 Schematics <strong>of</strong> generic ramjet engines.


30 FRY<br />

a) ATRJ<br />

b) ATR<br />

c) ERJ<br />

Fig. 4 Schematics <strong>of</strong> generic hybrid (mixed cycle) ramjets that produce<br />

static thrust.<br />

number variations are limited. The air-ducted rocket (ADR), shown<br />

in its IRR form in Fig. 3d, operates under the same engine cycle<br />

principles. Here, a fuel-rich monopropellant is used to generate a<br />

low-to-moderate pressure gaseous fuel supply for the subsonic combustor.<br />

The choice <strong>of</strong> an ADR is generally a compromise between<br />

the fuel supply simplicity <strong>of</strong> an SFIRR and the unlimited throttleability<br />

<strong>of</strong> the LFIRR, GFRJ, or CRJ. An ADR is generally used<br />

when the total fuel impulse does not adversely impact the powered<br />

range. The performance <strong>of</strong> the LFIRR or GFRJ is typically superior<br />

to the other four variants shown.<br />

Although the ramjets shown conceptually in Figure 3 are viable<br />

vehicle propulsion systems, none can produce static thrust. Figure 4<br />

shows three types <strong>of</strong> hybrid or combined cycle ramjet engine cycles<br />

that can. The first embeds a turbojet engine within the main ramjet<br />

engine and is usually liquid fueled and called an air-turboramjet<br />

(ATRJ) (Fig. 4a). Here, the turbojet produces the required static and<br />

low-speed thrust for take<strong>of</strong>f (and landing if required) that may or<br />

may not be isolated from the main ramjet flow at supersonic speeds.<br />

An alternative to the ATRJ is the air-turborocket (ATR) in which<br />

alow-to-moderate pressure rocket motor is used to drive a turbine<br />

and to provide a gaseous fuel for the ramjet (Fig. 4b). The turbine, in<br />

turn, drives a compressor, the combination <strong>of</strong> which produces static<br />

thrust. At supersonic speeds, the compressor, again, may be isolated<br />

from the main ramjet flow and the turbine idled so that the vehicle<br />

can operate as an ADR. The final hybrid ramjet cycle capable <strong>of</strong><br />

producing thrust is ejector ramjet (ERJ) shown in Fig. 4c. Here, a<br />

rocket motor or gas generator produces a high-pressure, generally<br />

fuel-rich, supersonic primary or ejector flow that induces secondary<br />

air to flow through the engine even at static conditions. The ejector<br />

effluent and air then mix and burn (at globally subsonic speeds) and<br />

finally expand in the convergent–divergent exit nozzle.<br />

There are three basic types <strong>of</strong> integral/rocket ramjet engines,<br />

namely, the LFRJ, the solid-fueled ramjet (SFRJ), and the ducted<br />

rocket (DR) as shown in Fig. 3. In the LFRJ, hydrocarbon fuel is<br />

injected in the inlet duct ahead <strong>of</strong> the flameholder or just before<br />

entering the dump combustor. In the SFRJ, solid ramjet fuel is cast<br />

along the outer wall <strong>of</strong> the combustor with solid rocket propellant<br />

cast on a barrier that separates it from the solid ramjet fuel. In this<br />

case, fuel injection by ablation is coupled to the combustion process.<br />

An aft mixer or other aid is generally used to obtain good<br />

combustion efficiency. The DR is really a solid fuel gas generator<br />

ramjet in which high-temperature, fuel-rich gasses are supplied to<br />

the combustor section. This provides for flame piloting and utilizes<br />

the momentum <strong>of</strong> the primary fuel for mixing and increasing to-<br />

tal pressure recovery. Performance <strong>of</strong> the DR may be improved by<br />

adding a throttle valve to the primary fuel jet, thereby allowing larger<br />

turndown ratios. This is known as a variable flow DR (VFDR). The<br />

reader is directed to Zucrow, 6 Dugger, 7 and others 8 for additional<br />

ramjet fundamentals.<br />

Supersonic Combustion (Scramjet) Cycle Behavior<br />

Turning now to supersonic combustion engines, Fig. 5 shows a<br />

generic scramjet engine and two hybrid variants. Figure 5a shows a<br />

traditional scramjet engine wherein air at supersonic or hypersonic<br />

speeds is diffused to a lower, albeit still supersonic, speed at station 4.<br />

Fuel (either liquid or gas) is then injected from the walls (holes, slots,<br />

pylons, etc.) and/or in-stream protuberances (struts, tubes, pylons,<br />

etc.), where it mixes and burns with air in a generally diverging<br />

area combustor. Unlike the subsonic combustion ramjet’s terminal<br />

normal shock system, the combined effects <strong>of</strong> heat addition and<br />

diverging area in the scramjet’s combustor, plus the absence <strong>of</strong> a<br />

geometric exit nozzle throat, generate a shock train located at and<br />

upstream <strong>of</strong> the combustor entrance, which may vary in strength<br />

between the equivalent <strong>of</strong> a normal and no shock. The strength <strong>of</strong> this<br />

shock system depends on the flight conditions, inlet compression or<br />

inlet exit Mach number M4, overall engine fuel/air ratio ER0, and<br />

supersonic combustor area ratio (A5/A4).<br />

The unique combination <strong>of</strong> heat addition in a supersonic airstream<br />

with a variable strength shock system plus the absence <strong>of</strong> a geometric<br />

throat permits the scramjet to operate efficiently over a wide range<br />

<strong>of</strong> flight conditions, that is, as a nozzleless subsonic combustion<br />

ramjet at low flight Mach numbers, M0 = 3–6, and as a supersonic<br />

combustion ramjet at higher flight Mach numbers, M0 > 5. At low<br />

M0 and high ER, the combustion process generates the equivalent <strong>of</strong><br />

a normal shock system and is initially subsonic, similar to that <strong>of</strong> a<br />

conventional subsonic combustion ramjet, but accelerates to a sonic<br />

or supersonic speed before exiting the diverging area combustor,<br />

which eliminates the requirement for a geometric throat. As ER<br />

decreases at this same M0, the strength <strong>of</strong> the precombustion shock<br />

system will also decrease to the equivalent <strong>of</strong> a weak oblique shock,<br />

and the combustion process is entirely supersonic. At high M0, the<br />

strength <strong>of</strong> the shock system is always equivalent to either a weak<br />

oblique shock or no shock, regardless <strong>of</strong> ER. This is referred to as<br />

dual-mode combustion and permits efficient operation <strong>of</strong> the engine<br />

from M0 = 3toM0 = 8–10 for liquid hydrocarbon fuels and up to<br />

a) Supersonic combustion ramjet<br />

b) DCR<br />

c) ESJ<br />

Fig. 5 Schematics <strong>of</strong> generic supersonic combustion engines.


orbital speeds for gaseous (hydrogen) fuels. The upper limit for<br />

the liquid-fueled cycle is, <strong>of</strong> course, due to energy consumption<br />

by dissociating and ionizing species at elevated temperatures that<br />

cannot be compensated for by additional fuel as in the case <strong>of</strong>, for<br />

example, a diatomic gas such as hydrogen.<br />

The supersonic combustion ramjet (scramjet) engine (Fig. 5) is<br />

ahypersonic airbreathing engine in which heat addition, due to the<br />

combustion <strong>of</strong> fuel and air, occurs in a flow that is supersonic relative<br />

to the engine. The potential performance <strong>of</strong> scramjet engines using<br />

hydrogen fuel and variable geometry covers a wide Mach number<br />

range, from M0 = 4to15+. Scramjet missiles using hydrocarbon<br />

fuels are usually <strong>of</strong> fixed geometry, except for inlet starting provisions,<br />

to minimize weight and complexity. Variable geometry scramjets<br />

can improve performance somewhat and are more suitable for<br />

vehicle applications where a broad Mach number range is needed.<br />

Both axisymmetric and two-dimensional scramjets have been developed.<br />

Axisymmetric engines are usually lighter weight but are<br />

more representative <strong>of</strong> high drag pod-mounted engines, whereas<br />

two-dimensional engines can be more easily integrated into a vehicle<br />

body.<br />

Unlike a conventional ramjet engine, where the incoming airflow<br />

is decelerated to a subsonic speed by means <strong>of</strong> a multishock intake<br />

system, the flow in a scramjet is allowed to remain supersonic.<br />

In this case, the amount <strong>of</strong> compression performed by the inlet<br />

can be significantly reduced and normal shock losses eliminated<br />

with a corresponding increase in total pressure recovery. This can<br />

more than compensate for the high heat addition losses (Rayleigh<br />

losses) encountered. In addition, the reduced level <strong>of</strong> compression<br />

results in lower static temperatures and pressures at the entry to the<br />

combustor, which reduces the severity <strong>of</strong> the structural loads. The<br />

reduced temperature allows more complete chemical reaction in<br />

the combustor and can reduce the losses due to finite-rate chemical<br />

reactions in the nozzle. 9 In reality, flow in the scramjet combustor<br />

can be mixed flow at this Mach number with regions <strong>of</strong> subsonic<br />

flow near surfaces and supersonic flow in the core.<br />

Two attractive approaches for providing (a broader Mach number<br />

operational range) scramjets with a lower Mach number capability<br />

are the dual-mode scramjet (DMSJ) engine in which both<br />

subsonic and supersonic modes <strong>of</strong> combustion are possible within<br />

one combustor and the hypersonic dual-combustor ramjet (DCR)<br />

engine. The dual-mode supersonic combustion ramjet engine was<br />

proposed in the early 1960s, U.S. Patent 3,667,233 by Curran and<br />

Stull, and was subsequently developed by the Marquardt Corporation<br />

(Marquardt) in their early DMSJ Engine. James Keirsey <strong>of</strong><br />

Johns Hopkins University/Applied Physics Laboratory (JHU/APL)<br />

invented the DCR cycle in the early 1970s.<br />

Characterization <strong>of</strong> the complex flow-field in DMSJs (first introduced<br />

by Curran and Stull in 1963) has been the subject <strong>of</strong> a number<br />

<strong>of</strong> previous investigations. Billig and Dugger, 10 Billig et al., 11<br />

and Waltrup and Billig 12,13 first provided analysis <strong>of</strong> experimental<br />

data and analytic tools, which allow a prediction <strong>of</strong> the flowfield<br />

such as upstream interaction and required isolator length for midspeed<br />

scramjet combustor configurations. A well-known correlation<br />

for upstream interaction distance was formulated with dependence<br />

specified in terms <strong>of</strong> heat release (downstream pressure rise) and<br />

the entering momentum characteristics <strong>of</strong> the boundary layer before<br />

isolator separation. Heiser and Pratt 14 provide a thorough and<br />

extended treatment <strong>of</strong> dual-mode flowfields.<br />

Although the scramjet <strong>of</strong>fers these unique capabilities, it also requires<br />

special fuels or fuel preparation to operate efficiently above<br />

M0 = 6 because <strong>of</strong> low static temperatures and short combustor<br />

residence times (


32 FRY<br />

Fig. 6 Characteristic performance by engine type.<br />

Fig. 7 Engine options as function <strong>of</strong> Mach number.<br />

over the first forty years and supplemented his work in 1999. 21<br />

There has been a considerable amount <strong>of</strong> laboratory testing and<br />

ground testing <strong>of</strong> scramjet engines in various countries, described<br />

in the cited references and briefly reviewed within this paper. A basic<br />

understanding <strong>of</strong> the supersonic combustion process has improved<br />

significantly over the years.<br />

Selection <strong>of</strong> engines for high speed vehicles is dependent on the<br />

intended mission and the speed range <strong>of</strong> interest. Figure 7 shows<br />

a summary <strong>of</strong> various options as a function <strong>of</strong> Mach number. The<br />

choice <strong>of</strong> propulsion options and the operating range can be broadened<br />

by combining different propulsion options to create combined<br />

cycle or combination cycle engines, as discussed in the preceding<br />

sections.<br />

Rationale favoring a Mach 6–8 missile includes providing substantially<br />

improved capabilities over a Mach 4 missile in terms <strong>of</strong><br />

ranges attained within a given flight time or less flight time for<br />

agiven range with the accompanying improved survivability. The<br />

higher kinetic energy and speed are additionally essential for increasing<br />

the probability <strong>of</strong> neutralizing various target types.<br />

Having introduced the ramjet and scramjet cycles, let us now<br />

return to the evolution <strong>of</strong> ramjet engines for powered flight.<br />

History <strong>of</strong> <strong>Ramjet</strong> and Scramjet Development<br />

Subsonic Combustion History<br />

Many excellent historical accounts chronicle the early years <strong>of</strong><br />

ramjet development. 2,5,20,22,23,35,36 Highlights <strong>of</strong> this history are reviewed<br />

here. Avery2 was among the first to review progress from<br />

the beginning years, while focusing on a 25-year period from the<br />

early 1930s, when testing <strong>of</strong> practical designs first began in 1955.<br />

Waltrup5 provided a very thorough review <strong>of</strong> international airbreathing<br />

engine development from the beginning years to 1987, with<br />

an emphasis on the development <strong>of</strong> supersonic combustion ramjet<br />

(scramjet) engines. In 1996, Waltrup et al. provided a history <strong>of</strong> U.S.<br />

Navy ramjet, scramjet and mixed-cycle propulsion developments. 22<br />

Kuentzmann and Falempin provide an account <strong>of</strong> French and international<br />

ramjet and scramjet development history. 24,34<br />

The history <strong>of</strong> ramjet concepts began in the early 1900s, with<br />

actual testing not beginning for another 30 years. While many international<br />

researchers were focusing on solving the thrust to weight<br />

challenges <strong>of</strong> internal combustion engines for flight, Lake (in the<br />

United States) and Lorin (in France) and their colleagues were examining<br />

jet propulsion devices that did not have in-stream obstructions,<br />

currently called ejector ramjets. The first patent <strong>of</strong> a subsonic<br />

ramjet cycle device, an ejector ramjet, was issued to Lake<br />

in 1909. Lorin published the first treatise on subsonic ramjets in<br />

1913, but did not address high subsonic or supersonic flight. Morize<br />

(France, 1917) and Melot (France, 1920) engineered the ejector ramjet<br />

concept, with testing occurring in France during World War I<br />

(WWI) and in the United States in 1927. Although tests demonstrated<br />

an increase in static thrust, interest in this engine cycle<br />

waned until the late 1950s. Carter (in Great Britain) patented the<br />

first practical ramjetlike device for enhancing the range <strong>of</strong> artillery<br />

shells in 1926. Carter’s designs showed considerable insight for<br />

the time and employed a normal shock inlet with either a conical<br />

nose/annular duct or central cylindrical duct. The first recognizable<br />

conical-nosed LFRJ patent was given to Fono (in Hungary) in 1928.<br />

These designs included convergent–divergent inlet diffuser, fuel injectors,<br />

flameholders, combustor, and convergent–divergent nozzle.<br />

Although these systems were designed for supersonic, high-altitude<br />

aircraft flight, they did not advance beyond the design stage.<br />

Actual construction and testing <strong>of</strong> viable ramjet designs did not<br />

occur until the mid-1930s in France, Germany, and Russia. Leduc<br />

(France) ground tested a conical ramjet up to Mach 0.9. Work on<br />

a full-scale ramjet-powered aircraft had begun by 1938, with component<br />

ground tests conducted up to Mach 2.35 by 1939. World<br />

War II(WWII) temporarily halted further testing. In Germany,<br />

Trommsdorff led a successful effort in 1935 to develop artillery<br />

shells powered by multiple-shock, conical-inlet LFRJs. These shells<br />

actually accelerated from Mach 2.9 to 4.2 in tests conducted in<br />

1940. Sänger and colleagues (in Germany) examined designs for<br />

an aircraft-launched ramjet-powered cruise missile but never constructed<br />

or tested one. The Germans did field the first operational<br />

ramjet-powered missile in the form <strong>of</strong> the V-1 buzzbomb powered<br />

by a subsonic flight speed pulsejet engine. Strechkin (in Russia) also<br />

began ground testing <strong>of</strong> ramjet components at speeds up to Mach 2 in<br />

the 1930s. Under Merkulov’s direction, Russia successfully flight<br />

tested a tandem-boosted ADR using magnesium/aluminum solid<br />

fuel in 1939. These activities were subsequently replaced with designs<br />

to augment the thrust <strong>of</strong> existing aircraft using wing-mounted<br />

ramjet pods. Attempts were also thwarted by the events <strong>of</strong> WWII<br />

following initial flight testing in 1940. Reid (in the United States)<br />

and Marquardt (in Great Britain) began ramjet development efforts<br />

in the early 1940s in the form <strong>of</strong> aerial-guided projectiles and aircraft<br />

performance augmenters, respectively. These efforts continued after<br />

WWII and resulted in weapon systems such as the Bomarc (U.S. Air<br />

Force), Talos (U.S. Navy), and Bloodhound (Great Britain) antiaircraft<br />

missiles, as well as numerous basic and applied experiments<br />

at national research centers in both countries.<br />

The ramjet engine began to receive attention during the secondhalf<br />

<strong>of</strong> the 1940s and reached a relative peak during the 1950s with<br />

a number <strong>of</strong> operational systems being deployed. France developed<br />

several operational ramjet missiles (VEGA, CT-41, and SE 4400)<br />

in the late 1950s and early 1960s. The ramjet, always needing an<br />

auxiliary propulsion system for starting, got squeezed between improved<br />

turbine engines and rockets during the 1950s and did not<br />

recover until reignited by the Russian SA-4, SA-6 and SS-N-19<br />

design revolutions in the late 1960s and early 1970s. This period<br />

witnessed a surge in development activity by the United States<br />

and Russia in the development <strong>of</strong> low-volume IRR missile designs.<br />

The Russian activity led to the operational missiles SS-N-22 and<br />

AS-17/Kh-31.<br />

The ramjet received expanded international development attention<br />

beginning in the 1980s, which continues today. During<br />

the 1980s France developed the operational Air-So Moyenne<br />

Portee-Ameliore (ASMP) and flight-tested Missile Probatoire Stato


Rustique (MPSR)/Rustique. The United States invested effort in supersonic<br />

low-altitude target (SLAT) and VFDR. During the 1990s,<br />

France continued its long history <strong>of</strong> development activity in ramjets<br />

with activity on MARS, MPSR/Rustique, Anti-Naivre Futur/Anti-<br />

Navire Nouvelle Generation ANF/ANNG, Vesta, and the nextgeneration<br />

ASMP-A. The People’s Republic <strong>of</strong> China began development<br />

<strong>of</strong> a long-range antiship variant (C-301) <strong>of</strong> its C-101 and<br />

the more advanced Hsiung Feng. South Africa began development<br />

<strong>of</strong> a long-range air-to-air missile (LRAAM). Russia continued to<br />

demonstrate its understanding <strong>of</strong> this technology by beginning the<br />

development <strong>of</strong> AA-X-12 and SS-N-26. Israel entered the ramjet<br />

community by beginning development <strong>of</strong> a ramjet-powered version<br />

<strong>of</strong> Gabriel for extended range. Germany began development <strong>of</strong> an<br />

antiradiation missile ARMINGER. India began development <strong>of</strong> the<br />

PJ-10/Brahmos, a derivative <strong>of</strong> the Russian SS-N-26. The 2000s<br />

have seen this development activity continue and expand yet further<br />

in the United States supersonic sea skimming target (SSST), generic<br />

supersonic cruise missile (GSSCM), and high-speed antirachation<br />

demonstration (HSAD), The United Kingdom beyond visual range<br />

air-to-air missile (BVRAAM/Meteor), France (MICA/RJ), and elsewhere.<br />

These and other international development programs are reviewed<br />

in further detail later in this paper as a means <strong>of</strong> better<br />

understanding the evolution in ramjet technology.<br />

Supersonic Combustion History<br />

Many excellent historical accounts chronicle international scramjet<br />

development. Curran provided a review <strong>of</strong> the first 40 years <strong>of</strong><br />

international scramjet engine development in 1997. 20 Waltrup reviewed<br />

international supersonic combustion development in 1987, 5<br />

and Waltrup et al. reviewed U.S. Navy scramjet and mixed-cycle engine<br />

development in 1996. 22 Van Wie 25 chronicled the 59-year history<br />

<strong>of</strong> JHU/APL contributions to the development <strong>of</strong> high-speed<br />

vehicles, highlighting five great APL propulsion pioneers, Avery,<br />

Dugger, Keirsey, Billig, and Waltrup in 2003. Andrews 26 provided<br />

avery thorough historical review <strong>of</strong> scramjet development and testing<br />

in the United States in 2001. McClinton et al. 28 provided a<br />

worthy review <strong>of</strong> engine development in the United States for space<br />

access applications in 2001. Escher 27 provided an excellent review<br />

<strong>of</strong> U.S. developments in combined airbreathing/rocket propulsion<br />

for advanced aerospace applications in 1999.<br />

Efficient airbreathing engines for operation into the hypersonic<br />

speed regime have been studied for over 40 years. The heart <strong>of</strong> these<br />

engine systems is the supersonic combustion ramjet (scramjet) cycle.<br />

Scramjet engine concept development, test facilities and instrumentation<br />

development, analysis method refinement, and component<br />

and engine testing have been pursued continually since the early<br />

1960s. Efforts to integrate the scramjet with higher and lower speed<br />

propulsion devices for space access have been investigated sporadically<br />

since the 1960s and continuously since 1984. McClinton et al. 28<br />

reviewed U.S. hypersonic airbreathing launch vehicle propulsion<br />

development efforts. The review addresses experiences and major<br />

accomplishments <strong>of</strong> historic programs; the goals, coordination between,<br />

and status <strong>of</strong> current programs in 2001; and a view <strong>of</strong> the<br />

future <strong>of</strong> hypersonic airbreathing propulsion development in the<br />

United States for future launch vehicles.<br />

In 2001 McClinton et al. 28 discussed scramjet development in<br />

the United States in terms <strong>of</strong> generations, first generation 1960–<br />

1973, second generation 1969–1984, third generation 1984–1994,<br />

and fourth generation from 1995-today. Each generation was<br />

distinguished by its unique contributions <strong>of</strong> the level <strong>of</strong> understanding<br />

<strong>of</strong> supersonic combustion.<br />

First Generation Scramjet Development (Beginning: 1960–1973)<br />

The origins <strong>of</strong> employing combustion in supersonic flows in the<br />

United States can be traced back to interest in burning fuels in external<br />

streams to either reduce the base drag <strong>of</strong> supersonic projectiles<br />

or to produce lift and/or thrust on supersonic and hypersonic airfoils<br />

in the early 1950s. 5 In Europe, interest in supersonic combustion<br />

paralleled that in the United States throughout the 1960s and<br />

1970s. The most extensive <strong>of</strong> the early (first generation) scramjet<br />

development programs in the United States was the NASA hyper-<br />

FRY 33<br />

sonic research engine (HRE) program. The HRE program, started<br />

in 1964, was crafted to develop and demonstrate flight-weight,<br />

variable-geometry, hydrogen-fueled and- cooled scramjet engine<br />

scramjet technology. In France, both fundamental and applied research<br />

was being pursued, with initial connected pipe testing being<br />

conducted by ONERA at Mach 6 conditions in the early 1970s. In<br />

Germany, most <strong>of</strong> the reported work was on supersonic combustion<br />

was <strong>of</strong>amore fundamental nature. Russia had an extensive program<br />

in supersonic combustion and scramjet propulsion since the<br />

1960s. Canadian interest in supersonic combustion began in 1960<br />

at MacGill University in hypersonic inlet aerodynamics and gunlaunched<br />

scramjet flight testing.<br />

The first Generation witnessed the start <strong>of</strong> several major scramjet<br />

development programs in the U.S. during the mid-1960s, following<br />

the first scramjet demonstration by Ferri in 1960 29 and studies<br />

that verified the benefits <strong>of</strong> scramjet propulsion. The U.S. Air<br />

Force, NASA, and U.S. Navy sponsored programs tested six scramjet<br />

engines/flowpaths through major contracts at Marquardt, General<br />

Electric, United <strong>Technology</strong> Research Center (UTRC), Garrett, and<br />

General Applied Sciences Laboratory (GASL). Engine flowpaths<br />

from all <strong>of</strong> these contractors were tested at low hypersonic speeds<br />

(up to Mach 7). Most tests utilized hydrogen fuel, but the U.S. Air<br />

Force also funded hydrocarbon-fueled scramjet tests for missile applications,<br />

and the U.S. Navy (APL) performed several different hydrocarbon<br />

studies in the Supersonic Combustion <strong>Ramjet</strong> (SCRAM)<br />

project. However, the U.S. Air Force withdrew from scramjet research<br />

and development, not to return until the National Aerospace<br />

Plane (NASP) program in 1984. Following successful demonstration<br />

<strong>of</strong> scramjet performance, operability and structural/systems integration,<br />

NASA turned to development and validation <strong>of</strong> airframeintegrated<br />

engine flowpaths.<br />

Second Generation Scramjet Development (Airframe Integration:<br />

1973–1986)<br />

The technology development focus in the United States shifted<br />

in the second generation to integration <strong>of</strong> hydrogen-fueled scramjet<br />

engines on a hypersonic vehicle following the first generation<br />

hypersonic propulsion demonstrators. A Mach 7 cruiser configuration<br />

was selected with turbojet low-speed and scramjet high-speed<br />

systems, in an over–under arrangement. NASA Langley Research<br />

Center (LaRC) led this effort, which focused on the sidewall compression<br />

scramjet engine. This engine included fuel injector struts<br />

and fixed geometry to minimize weight. Much <strong>of</strong> the research effort<br />

was placed on tool development. This included facilities, test<br />

methodologies, cycle analysis, data analysis, and computational<br />

fluid dynamics, (CFD). This time frame was also characterized by<br />

adownturn in research in the United States, and facility availability<br />

became an issue. Therefore, facilities were developed at NASA<br />

LaRC for efficient scramjet testing. Modest-sized facilities also remained<br />

available at GASL and were used to handle higher pressure<br />

validation tests. Component test facilities were also developed, including<br />

a combustion-heated direct-connect combustor facility and<br />

a Mach 4, high-pressure inlet test facility.<br />

Component tests were performed to establish rules-<strong>of</strong>-thumb<br />

design models/tools. These models 30 were not incorporated into<br />

the U.S. scramjet toolbox until the late 1980s. Component tests<br />

were also used to develop databases for verification <strong>of</strong> analytical<br />

and computational methods. These second generation cycle analysis<br />

methods are simplistic by today’s standards <strong>of</strong> CFD methods,<br />

nevertheless feature internal calculations such as shear and heat<br />

transfer to the combustor wall, fuel mixing, and estimated finite-rate<br />

chemistry effects on combustion. Scramjet tests for the three-strut<br />

engine were performed at NASA LaRC and GASL.<br />

Aerodynamic and propulsion-airframe integration (PAI) assessment<br />

is required to set scramjet performance requirements. Aerodynamic<br />

and PAI tests were performed to quantify inlet capture,<br />

external nozzle performance, and scramjet-powered vehicle performance.<br />

Aerodynamic wind-tunnel tests were performed at flight<br />

conditions up to Mach 8. Structural designs for the sidewall compression<br />

engine were developed, including primary structure, cooling<br />

jackets, and thermal management. These designs used cooling


34 FRY<br />

channels rather than the <strong>of</strong>fset fin approach used in the HRE, but<br />

maintained the high-temperature steel. France launched the ESOPE<br />

program, inspired at least in part by NASA’s HRE activity.<br />

During the mid-1970s, the sidewall compression flowpath tests<br />

demonstrated the required thrust, operability, and fuel cooling requirements<br />

to allow a credible vehicle design. About 1000 tests were<br />

performed on three engines. In addition, these studies validated the<br />

predicted scramjet performance and provided some justification for<br />

starting the NASP program.<br />

Scramjet module and direct-connect research and testing using<br />

gaseous hydrocarbon fuels was started at NASA LaRC in late 1970s<br />

and was subsequently interrupted by the NASP program. After<br />

NASP, the U.S. Air Force took the lead in this area. Tests were<br />

performed using methane, ethane, and ethylene injected from the<br />

hydrogen fuel injectors.<br />

Third Generation Scramjet Development (NASP: 1986–1994)<br />

In the early 1980s the U.S. NASP program was formulated,<br />

with the objective <strong>of</strong> developing a single-stage-to-orbit “hypersonic<br />

combined-cycle airbreathing capable” 31 engine to propel a research<br />

vehicle, the X-30. The NASP program promise <strong>of</strong> flying a singlestage<br />

vehicle, powered by a combined cycle engine that utilized<br />

scramjet operation to Mach 25 was aggressive, when the state <strong>of</strong><br />

technology in 1984 is considered. Subsequent development activity<br />

backed <strong>of</strong>f such an aggressive approach. Neither scramjet engines<br />

nor flowpaths had been tested above Mach 7. In addition, no credible,<br />

detailed analysis <strong>of</strong> scramjet performance, operability, loads,<br />

or structural approaches had ever been performed for flight past<br />

Mach 7. Also, what was good enough for Mach 7 vehicle operation<br />

was not refined enough for the low-thrust margin (energy from<br />

combustion vis-à-vis air kinetic energy) at double-digit flight Mach<br />

numbers. 30 In other words, second generation scramjet technology<br />

wasagood starting point, but considerable refinement and development<br />

was needed.<br />

International activity in this period included many developments.<br />

Germany began development <strong>of</strong> Sänger II in the late 1980s as a proposed<br />

two-stage-to-orbit (TSTO) concept vehicle. Japan pursued development<br />

<strong>of</strong> combined cycle engine technology for flyback booster<br />

TSTO applications. Russia developed Kholod as a first generation<br />

hypersonic flying laboratory, derived from the SA-5 (S-200 family)<br />

surface-to-air missile. Russia initiated the comprehensive hypersonic<br />

research and development in the ORYOL program with the<br />

purpose <strong>of</strong> developing combined propulsion systems for advanced<br />

reusable space transportation. Finally, Russia employed another first<br />

generation flight-test vehicle GELA Phase I testbed for the development<br />

<strong>of</strong> Mach 3 ramjet missile propulsion systems. France initiated<br />

PREPHA aimed at developing a knowledge base on hydrogenfueled<br />

dual-mode ramjet technology for single-stage-to-orbit applications.<br />

Fourth Generation Scramjet Development (Resurgence: 1995–Today)<br />

Following the NASP program, three new directions were taken<br />

in the United States. The U.S. Air Force went back to hydrocarbonfueled<br />

scramjet missiles; NASA aeronautics went on to demonstrate<br />

the most advanced parts <strong>of</strong> the NASP propulsion technology,<br />

that is, scramjets; and the NASA rocket community embraced the<br />

engine technology afforded by rocket-airbreathing combined cycle<br />

engines. These three programs, HyTech/HySet, Hyper-X, and<br />

Spaceliner, are mentioned here, and their program contributions<br />

and status are reviewed later. The United States has since incorporated<br />

the development <strong>of</strong> high-speed airbreathing technology<br />

within an overarching approach called the National Aerospace Initiative<br />

(NAI). It is a partnership between the Department <strong>of</strong> Defense<br />

(DOD) and NASA designed to sustain the U.S. leadership<br />

through technology development and demonstration in three pillar<br />

areas <strong>of</strong> high speed/hypersonics, space access, and space technology.<br />

Ronald Sega, Director <strong>of</strong> the U.S. Defense Research and<br />

Engineering Agency (DARPA), points out that NAI will provide<br />

many benefits: never-before-available military capabilities to satisfy<br />

a broad range <strong>of</strong> needs; technologies required to provide reliable<br />

and affordable space transportation for the future, develop<br />

launch systems, and satisfy exploratory mission needs; and, finally,<br />

spur innovation in critical technology areas and excite and inspire<br />

the next-generation high-technology science and engineering workforce<br />

in the United States.<br />

HyTech/HySET. The goal <strong>of</strong> the U.S. Air Force Hypersonic<br />

<strong>Technology</strong>/Hydrocarbon Scramjet Engineering <strong>Technology</strong><br />

(HyTech/HySET) program is to advance technology for liquid<br />

hydrocarbon-fueled scramjets. Although this technology will be applicable<br />

to scramjet-powered strike, reconnaissance, and space access<br />

missions, the initial focus is on missile scale and applications.<br />

The HyTech/HySET program has made significant advancements<br />

over the past 8 years in the following issues associated with<br />

liquid hydrocarbon-fueled scramjet engine development: ignition<br />

and flameholding methodologies, endothermic fuels technology,<br />

high-temperature materials, low-cost manufacturing technology for<br />

scramjet engines, and detection and cleaning procedure for coked<br />

heat exchangers. The engine development addressed issues associated<br />

with weight, cost, and complexity. An effective fixed-geometry<br />

scramjet engine was developed for operation over the Mach 4–8<br />

speed range. HySET was unique in having developed scramjet performance<br />

and structural durability <strong>of</strong> complete engine configurations<br />

not just flowpaths.<br />

Hyper-X. NASA initiated the joint LaRC and Dryden Flight<br />

Research Center hypersonic X-plane (Hyper-X) program in 1996 to<br />

advance hypersonic airbreathing propulsion (scramjet) and related<br />

technologies from the laboratory to the flight environment. This<br />

is to be accomplished using three small (12-ft long), hydrogenfueled<br />

research vehicles (X-43) flying at Mach 7 and 10. The Hyper-<br />

X program technology focus is on four main objectives required<br />

for practical hypersonic flight: Hyper-X (X-43) vehicle design and<br />

flight-test risk reduction, flight validation <strong>of</strong> design methods, design<br />

methods enhancement, and Hyper-X phase 2 and beyond.<br />

Hyper-X Phase 2 and beyond activities 32 include program planning,<br />

long-term, high-risk research, and refinement <strong>of</strong> vision vehicle<br />

designs. <strong>Propulsion</strong> related development activity in this arena<br />

includes the evaluation <strong>of</strong> the pulse detonation engine (PDE) for<br />

hypersonic systems, magnetohydrodynamics (MHD) scramjet studies,<br />

and design developments leading to highly variable-geometry<br />

scramjets. Powered take<strong>of</strong>f and landing and low-speed operation <strong>of</strong><br />

ahypersonic shaped vehicle using remotely piloted vehicles will<br />

address the low-speed PAI issue identified in NASP. Finally, this<br />

arena was active in planning/advocating future directions for space<br />

access vehicle and airbreathing propulsion development. 19<br />

Third Generation Space Access. During the late 1990’s NASA<br />

established long term goals for access-to-space. NASA’s third generation<br />

launch systems are to be fully reusable and operational (IOC)<br />

by 2025. 33 The goals for third generation launch systems are to<br />

reduce cost by a factor <strong>of</strong> 100 and improve safety by a factor <strong>of</strong><br />

10,000 over current conditions. NASA’s Marshall Space Flight Center<br />

in Huntsville, AL has the agency lead to develop third generation<br />

space transportation technologies. Development <strong>of</strong> third generation<br />

launch vehicle technology falls under NASA’s Space Transportation<br />

Program. The programs have had names like Spaceliner, Advanced<br />

Space Transportation Program (ASTP), and the Hypersonic Investment<br />

Area <strong>of</strong> Next Generation Launch <strong>Technology</strong> (NGLT). These<br />

programs focus development <strong>of</strong> technologies in two main areas:<br />

propulsion and airframes. The program’s major investment is in<br />

hypersonic airbreathing propulsion since it <strong>of</strong>fers the greatest potential<br />

for meeting the third generation launch vehicle goals. The<br />

program is maturing the technologies in three key propulsion areas,<br />

scramjets, rocket-based combined cycle and turbine-based combination<br />

cycle. Ground and flight propulsion tests are underway or<br />

planned for the propulsion technologies. Airframe technologies are<br />

matured primarily through ground testing. Selection and prioritization<br />

<strong>of</strong> technology is guided by system analysis for third generation<br />

“vision” vehicles. These vehicles are generally two-stage-to-orbit


vehicles, which can be interrogated for safety, reliability and cost<br />

impacts <strong>of</strong> the proposed technologies.<br />

Flight-tests supporting NASA’s Third Generation Space Access<br />

focuses on incremental development and demonstration <strong>of</strong> key technology<br />

that can not be demonstrated to a <strong>Technology</strong> Readiness<br />

Level (TRL) <strong>of</strong> 6 in ground tests. These start with scramjet performance,<br />

operability and airframe integration (X-43A, X-43C and<br />

X-43D) for flight Mach numbers from 5 to 15. These first demonstrators<br />

are expendable to reduce test costs. The next step is integration<br />

<strong>of</strong> low-speed (Mach 0 to 3+) with the scramjet system in a reusable<br />

“combined cycle” flight demonstration (RCCFD). The first step for<br />

RCCFD is a Mach 0-7 reusable air-launched research vehicle, similar<br />

in size to the X-15. The final step would be a larger vehicle<br />

capable <strong>of</strong> operation to full airbreathing Mach number required for<br />

the 2025 IOC vehicle.<br />

International activity in this period continued to include many<br />

developments. A joint French/Russian program Wide Range <strong>Ramjet</strong><br />

(WRR) was initiated to develop technology for reusable<br />

space launcher applications. France also began the development<br />

<strong>of</strong> Joint Airbreathing Research for Hypersonic Application Research<br />

(JAPHAR) in cooperation with Germany as follow-on to<br />

the PREPHA (France) and Sänger (Germany) programs to pursue<br />

hypersonic airbreathing propulsion research for reusable space<br />

launcher applications. Furthermore France initiated Promethée<br />

aimed at developing fully variable-geometry endothermic hydrocarbon<br />

fuel dual-mode ramjet technology for military applications.<br />

The French efforts are leading toward LEA, a new flight-test demonstration<br />

<strong>of</strong> a high-speed dual-mode ramjet propelled vehicle at flight<br />

conditions <strong>of</strong> Mach 4–8 in the 2010–2012 time frame. In this era,<br />

Russia began openly discussing their development <strong>of</strong> AJAX, an<br />

innovative hypersonic vehicle concept envisioned to capture and recycle<br />

energy otherwise lost in flight at high Mach numbers. Russia<br />

also initiated second generation hypersonic flying laboratory work<br />

with Gela Phase II and the Mig-31 HFL. Australia conducted, with<br />

international support, the world’s first verified demonstration <strong>of</strong> supersonic<br />

combustion in a flight environment under HyShot.<br />

The development <strong>of</strong> vehicles for space access applications reflected<br />

maturation in propulsion technology and the technical interests<br />

<strong>of</strong> the times. Initial concepts, for example, the German<br />

Sänger–Bret Silbervögel <strong>of</strong> 1938, postulated single-stage-to-orbit<br />

(SSTO) vehicles based on pure rocket systems, or rocket-l<strong>of</strong>ted<br />

boost-gliders, such as the U.S. Dyna-Soar. The U.S. Air Force supported<br />

Spaceplane development followed, which spawned imaginative<br />

upper atmospheric high-Mach air collection and oxygen<br />

extraction technologies. The understanding <strong>of</strong> scramjet technology<br />

had began. By the early 1960s, the maturation <strong>of</strong> advanced<br />

airbreathing technology encouraged a redirection toward complex<br />

fully reusable TSTO vehicles with airbreathing first stages (with<br />

combinations <strong>of</strong> turbojets, turboramjets, or ramjets/scramjets) and<br />

rocket-boosted second stages. The economic realities <strong>of</strong> the 1970s<br />

dictated using semi-expendable, pure-rocket approaches, typified<br />

by the space shuttle. The potentialities <strong>of</strong> the advanced airbreathing<br />

scramjet <strong>of</strong> the 1980s led to NASP and horizontal take<strong>of</strong>f and landing<br />

concepts for airbreathing SSTO vehicles using complex propulsion<br />

systems dependent on new high-energy fuel concepts and the<br />

air collection and oxygen extraction technologies developed previously.<br />

The 1990s witnessed less ambitious goals <strong>of</strong> developing<br />

either pure advanced rocket systems (X-33 and X-34) or systems<br />

using straightforward scramjet technology (X-43A Hyper-X). The<br />

first flight demonstration <strong>of</strong> scramjet-propelled vehicle designs with<br />

true potential for enabling space access promises to become reality<br />

in 2003–2004.<br />

These and other international development programs are reviewed<br />

in further detail later as a means <strong>of</strong> better understanding the evolution<br />

in scramjet technology.<br />

<strong>Evolution</strong> in <strong>Ramjet</strong> and Scramjet<br />

Development Programs<br />

The development <strong>of</strong> ramjet technology for multiple applications<br />

has proceeded in parallel throughout history. Applications have<br />

ranged from boost- to main-propulsion for aircraft, gun projectiles,<br />

FRY 35<br />

missiles, and space launch vehicles. The intent here is to highlight<br />

international activity and selected programs as a means <strong>of</strong> identifying<br />

sources <strong>of</strong> technology advances, <strong>of</strong>ten resulting from parallel<br />

efforts in multiple applications.<br />

The key enabling technologies, components, or events that had<br />

significant impact on the maturation <strong>of</strong> ramjet propulsion and engine<br />

designs, summarized in Table 1, are briefly discussed relevant to the<br />

worldwide development <strong>of</strong> vehicle concepts and systems. The history<br />

<strong>of</strong> the worldwide subsonic and supersonic combustion ramjet<br />

evolution is summarized by era in Tables 2–6 from the turn <strong>of</strong> the<br />

century to today. Presented in Tables 2–6 for each ramjet/scramjet<br />

system are known historical era, originating country, engine/vehicle<br />

name, engine cycle type, development dates, design cruise Mach<br />

number, altitude and range performance, system physical characteristics,<br />

and state <strong>of</strong> development. The ranges provided are a mix<br />

depending on the engine/vehicle development status: operational<br />

range for operational systems, predicted range for concept vehicles,<br />

or demonstrated range for flight-test vehicles. Additionally, those<br />

engines/vehicles that are discussed and illustrated in this paper are<br />

indicated. Many observations can be drawn from the data shown in<br />

Tables 2–6 and include the following.<br />

1) <strong>Ramjet</strong> technology development has been consistently pursued<br />

internationally from very early days, accompanied by steady<br />

increases in airbreathing system capabilities.<br />

2) <strong>Ramjet</strong> engines have received substantially more attention than<br />

scramjet engines, with scramjet development increasing steadily<br />

since the early 1990s, which reflects the accelerating pace in the<br />

solution <strong>of</strong> the challenges <strong>of</strong> high speed flight.<br />

3) Although at the verge <strong>of</strong> success, flight testing <strong>of</strong> ramjetpowered<br />

engine concepts at hypersonic speeds has yet to be accomplished.<br />

At this writing the U.S. X-43A Hyper-X is planned for<br />

a second flight test in January 2004.<br />

<strong>Ramjet</strong> Development 1918–Today<br />

<strong>Ramjet</strong> Development: 1918–1960<br />

This era saw the birth <strong>of</strong> ramjet-powered aircraft flight and its<br />

rapid maturing <strong>of</strong> technology into primarily missile applications<br />

and its transition from subsonic to supersonic ramjet flight. Table 2<br />

summarizes ramjet evolution in the era from 1918 to 1960 and provides<br />

originating country, engine/vehicle name, development dates,<br />

performance, physical characteristics, and state <strong>of</strong> development.<br />

Countries actively engaged in development in this era were France,<br />

Germany, the United States, the United Kingdom and Russia.<br />

The Germans flew the first operational ramjet-powered missile<br />

in 1940 in the form <strong>of</strong> the V-1 buzzbomb (Fig. 8) powered by a<br />

subsonic flight speed pulsejet engine launched by a solid propellant<br />

booster. The V-1 could be considered the first cruise missile.<br />

German engineer, Paul Schmidt, working from a design <strong>of</strong> the Lorin<br />

tube, developed and patented (June 1932) a ramjet engine (Argus<br />

pulse jet) that was later modified and used in the V-1 Flying Bomb.<br />

The German V-1 technology was transferred to other countries after<br />

WWII.<br />

The first ramjet-powered airplane was conceived and variants<br />

tested by Leduc <strong>of</strong> France. The first powered flight <strong>of</strong> a ramjetpowered<br />

aircraft, Leduc-010 (Fig. 9), took place in April 1949<br />

Fig. 8 German V-1 operational WW II missile (1933–1945).


36 FRY<br />

Table 2 Worldwide ramjet evolution 1918–1960<br />

Powered Total Total<br />

Cruise Cruise range, length, Diameter, weight, State <strong>of</strong><br />

Era Country/service Engine/vehicle Engine type Dates, year Mach no. altitude, ft n mile Launcher in. in. lbm development<br />

1918–1945 France Melot research ERJ 1918–1927 0.4 —— —— —— —— —— —— Component tests<br />

1930–1955 France Leduc aircrafta Liquid-fueled 1933–1951 0.4–2.1 10,000 —— Ground —— —— —— Flight tests<br />

ramjet (LFRJ)<br />

Russia Stechkin research LFRJ 1933–1936 0.4–2.0 —— —— —— —— —— —— Component tests<br />

Germany V-1a Pulse jet 1933–1945 0.3 20,000 220 Rail 312 30 4,800 Operational<br />

Germany Sänger 1a <strong>Ramjet</strong> (RJ) 1936–1945 4+ 70,000 9,000 Rail 1,100 —— 200,000 Concept vehicle<br />

Russia Antipodal bomberb RJ 1945–1955 4+ 70,000 9,000 Rail 1,100 —— 200,000 Concept vehicle<br />

Russia Merkulov research LFRJ 1936–1939 0.3 —— —— —— —— —— —— Flight tests<br />

United States KDH-1 droneb Pulse jet 1942–1946 0.3 20,000 145 Rail 134 8 —— Operational<br />

1945–1960 United States Cobrab LFRJ 1945–1946 2.0 20,000 —— Rail —— 6 240 Flight tests<br />

United States Gorgon IV LFRJ 1945–1955 0.4 20,000 —— —— —— —— —— Flight tests<br />

United States KDM-1 Droneb LFRJ 1945–1949 2 35,000 60 Rail 264 30 1,600 Operational<br />

United States P-51/F-80/B-26 LFRJ 1945–1955 0.4 20,000 —— Air —— —— —— Flight tests<br />

United States NAVAHOa,g LFRJ 1946–1958 3.25 37,000–90,000 5,500 Ground 1,050 48 120,000 Flight tests<br />

c<br />

United States BTV LFRJ 1947–1948 2.4 30,000 10 Rail —— 18 —— Flight tests<br />

d<br />

U.S. Navy RTV LFRJ 1949–1950 2.4 30,000 25 Rail —— 24 —— Flight tests<br />

Russia Burya<br />

e<br />

b LFRJ 1950–1958 3.15 35,000–70,000 4,320 Ground 1,000 67 130,000 Flight tests<br />

U.S. Air Force BOMARCa,f LFRJ 1950–1972 2.5–3.4 70,000–100,000 440 Ground 560 35 15,620 Operational<br />

United States X-7a LFRJ 1950–1959 4.3 84,000 —— Rail —— 28,36 —— Flight tests<br />

U.S. Navy Talosa LFRJ 1950–1980 2.7 70,000 120 Rail 386 28 7,720 Operational<br />

France SIRIUS LFRJ 1950–1970 2.7 —— —— —— —— —— —— Component tests<br />

France Griffon IIb ATRJ 1951–1961 2.2 50,000 —— —— —— —— —— Flight tests<br />

U.S. Navy Trition/SSGM LFRJ 1951–1958 3.0 70,000 2,000+ Rail/ —— —— —— Component tests<br />

submarine<br />

U.S. Navy RARE SFIRR 1955–1960 2.3 50,000 —— Air 120 5 153 Flight tests<br />

France VEGAb b<br />

LFRJ 1955–1965 4.2 80,000 —— Rail —— —— —— Operational<br />

U.S. Navy CROWb SFIRR 1956–1964 3.0 50,000 97 Air 127 8 370 Flight tests<br />

France CT-41b LFRJ 1957–1965 3 75,000 —— Rail —— —— —— Operational<br />

France SE 4400b LFRJ 1957– 3.7 22,000 —— Rail —— —— —— Operational<br />

U.S. Navy Typhona LFRJ 1957–1965 4.1 100,000 200 Rail 333 16.75 6,160 Flight tests<br />

United Kingdom Bloodhounda LFRJ 1957–1991 2 23,000 108 Rail 306 21.5 2,300 Operational<br />

a b c d e f<br />

System discussed and shown. System discussed. Burner test vehicle (BTV). <strong>Ramjet</strong> test vehicle (RTV). Surface-to-surface quided vehicle (SSGM). Boeing Company and University <strong>of</strong> Michigan Aeronautics Research<br />

Center (BOMARC). g North American Aviation designation (NAVAHO).


Table 3 Worldwide ramjet evolution 1960–1980<br />

Powered Total Total<br />

Cruise Cruise range, length, Diameter, weight, State <strong>of</strong><br />

Era Country/service Engine/vehicle Engine type Dates, year Mach no. altitude, ft n mile Launcher in. in. lbm development<br />

FRY 37<br />

1960–1970 France Stataletexb LFRJ 1960–1970 3.8–5 25,000 —— Rail —— —— —— Flight tests<br />

U.S. Army Redhead Roadrunnerb LFRJ 1960–1980 2.5 0–60,000 217 Rail 298 12 900 Flight tests<br />

United Kingdom Sea Darta LFRJ 1960–1975 3.5 80,000 35 Rail 173 15.6 1,200 Operational<br />

United States Hyperjetb LFRJ 1960–1966 5+ 30,000 —— Air —— —— —— Flight tests<br />

United States D-21a LFRJ 1963–1980 4 1,00,000 2,600 Air 730 60 —— Operational<br />

Russia SA-4/Ganefa LFRJ 1964– 4 15,000 30 Rail 346 34 5,500 Operational<br />

U.S. Air Force LASRMb LFIRR 1964–1967 2.5 0–6,000 50 Air 168 15 1,566 Flight tests<br />

c<br />

U.S. Navy ATP/TARSAM-ER ADR 1965–1971 3.8 50,000–70,000 160 Rail 348 13.5 6,420 Component tests<br />

U.S. Navy ATP/TARSAM-MR<br />

h<br />

i<br />

j<br />

g<br />

c<br />

ADR-IRR 1965–1971 3.8 50,000–70,000 80 Rail 200 13.5 1,750 Component tests<br />

U.S. Air Force ASALMa LFIRR 1965–1980 2.5–4 10,000–80,000 56 Air 168 20 2,415 Flight tests<br />

Russia SA-6/Gainfula DRIRR 1965– 2.8 55,000 15 Rail 244 13.2 1,320 Operartional<br />

France SE X 422a LFRJ 1965–1967 2 30,000 —— Rail 394 20 4,190 Flight tests<br />

U.S. Navy IRR-SAM LFIRR 1966–1970 3.3 80,000 —— Rail 220 14.75 2,200 Component tests<br />

U.S. Navy ALVRJ-STMa LFIRR 1968–1979 3.0 0–40,000 28–100 Air 179 15 1,480 Flight tests<br />

1970–1980 U.S. Navy IRR-SSM LFIRR 1971–1974 2.5 50,000 —— Rail 200 14.75 2,000 Free-jet tests<br />

Russia SS-N-19/Shipwrecka LFRJ 1972– 2 0 340 Rail/sub 394 33 15,430 Operational<br />

U.S. Navy GORJEb LFIRR 1972–1976 2.6 0 35 Air 168 12 750 Semi-free-jet tests<br />

U.S. Navy ASAR d<br />

LFIRR 1972–1981 3.8 80,000 —— VLS 220 16 2,650 Semi-free-jet tests<br />

France Scorpiona LFRJ 1973–1974 6 70,000 325 Rail —— —— —— Component tests<br />

U.S. Navy MRE e<br />

LFIRR 1973–1977 3.0 30,000–70,000 150 Air 168 15 1,500 Free-jet tests<br />

Russia GELA phase Ia LFRJ 1973–1978 3 0–55,000 —— TU-95 —— —— —— Flight tests<br />

U.S. Navy IRR-TTV f /TTM f<br />

LFIRR 1974–1985 3.0 60,000 —— Submarine 246 21 3,930 Component tests<br />

U.S. Air Force ASALM–PTVa LFIRR 1976–1984 2.5–4 10,000–80,000 56 Air 168 20 2,415 Flight tests<br />

U.S. Air Force LFRED LFIRR 1976–1980 3 30,000 —— Air —— —— —— Component tests<br />

U.S. Navy LIFRAM LFIRR 1976–1980 3.0+ —— 150 Air 144 8 650 Semi-free-jet tests<br />

U.S. Navy SOFRAM SFIRR 1976–1981 3.0+ —— 150 Air 144 8 650 Free-jet tests<br />

Germany EFA SFDR 1976–1980 —— —— —— —— —— —— —— Flight tests<br />

U.S. Navy Firebrandb LFRJ 1977–1982 3 0 44 Rail 200 14 1,200 Flight test vehicle<br />

U.S. Air Force DRED DR 1977–1979 3 30,000 —— —— —— —— —— Component tests<br />

Russia SS-N-22/Sunburna LFIRR 1977– 2.5 0 270 Rail 360 28 8,800 Operational<br />

Russia AS-17/Kh-31a LFIRR 1977– 3 0 44–90 Rail 204 14 1,322 Operational<br />

France Rustique MPSR-1b UFDR 1978–1985 3 40,000 35 Rail 144 8 360 Flight tests<br />

People’s Republic C-101b LFRJ 1978– 1.8 0 30 Rail 295 21 4,070 Operational<br />

<strong>of</strong> China<br />

U.S. Army SPARK SFRJ/IRR 1978–1981 3 0 —— Rail —— —— —— Flight tests<br />

U.S. Air Force DR–PTV FFDR 1979–1986 3.5 60,000 —— Air —— —— —— Free-jet tests<br />

a b c d<br />

System discussed and shown. System discussed. Augmented thrust propulsion/thrust augmented surface-to-air missile-extended range, and -medium range (ATP/TARSAM-ER, -MR). Advanced surface-to-air ramjet<br />

(ASAR). e Modern ramjet engine (MRE). f Torpedo tube vehicle/torpedo tube missile (TTV/TTM). g Ducted rocket engine development (DRED). h Liquid fuel ramjet engine development (LERED). i Liquid fuel ramjet<br />

missile (LIFRAM). j Solid fuel ramjet missile (SOFRAM).


38 FRY<br />

Table 4 Worldwide ramjet evolution 1980–today<br />

Cruise Powered Total Total<br />

Cruise altitude, range, length, Diameter, weight, State <strong>of</strong><br />

Era Country/service Engine/vehicle Engine type Dates, year Mach no. ft n mile Launcher in. in. lbm development<br />

1980–1990 France/Germany ANS c<br />

LFIRR 1980–1992 2.5 0 44 Rail 204 14 1,322 Flight tests<br />

France ASMPa LFIRR 1980– 3 0 185 Air 210 15 1,848 Operational<br />

U.S. Navy ACIMD/AAAMa LFIRR 1981–1989 3.2 80,000 150 Air 144 9.5 590 Component tests<br />

U.S. Air Force BSFRJ d Tech SFRJ 1982–1987 3 70,000 300 Air —— —— —— Component tests<br />

U.S. Navy Vandal (Talos) a LFRJ 1983– 2.2 0 43.5 Rail 434 28 8,210 Operational<br />

U.S. Navy SFIRR SFIRR 1984–1989 2.5 0 50 Air 168 18 —— Component tests<br />

U.S. Navy SLATa LFIRR 1986–1992 2.5 0 50 Rail 218 21 2,445 Flight tests<br />

U.S. Air Force BECITE e<br />

SFIRR 1987–1990 3 70,000 300 Air —— —— —— Component tests<br />

U.S. Air Force VFDRa DR 1987–1997 3 50,000 50 Air 144 7 360 Component tests<br />

U.S./Germany Cooperative technical program SFDR 1989–1999 3 50,000 50 Air 144 7 360 Component tests<br />

1990–2000 France ASLP D2 LFIRR 1990–1996 2–3.5 —— —— Air —— 16 —— Component tests<br />

France MARSb LFIRR 1990–1996 4–5 90,000 —— Air 220 15 —— Component tests<br />

India AKASHb DRIRR 1990– 3.5 50,000 15 Rail 256 13 1,445 Flight tests<br />

United States/France Cooperative technical program UCDR 1991–1997 3 50,000 50 Air 144 7 360 Component tests<br />

People’s Republic <strong>of</strong> China C-301b LFRJ 1991– 2 0–20,000 110 Rail 388 30 10,100 Operational<br />

United States/Japan Cooperative technical program VFDR 1992–1997 3 50,000 50 Air 144 7 360 Component tests<br />

France Rustique MPSR-2b UFDR 1993–1997 3.5 80,000 30 Rail 158 7.9 440 Flight tests<br />

Russia MA-31/AS-17b LFIRR 1994– 4.5 0 44–90 Rail 200 14 1,400 Flight tests<br />

South Africa LRAAMb LFIRR 1994– 4 80,000 45 Air 144 7 360 Flight tests<br />

U.S. Navy LDRJ/LCMSb LFIRR 1995–1999 4.0 7,000 700 Air 256 21 3,400 Flight tests<br />

France/Germany ANF/ANNGb LFIRR 1995–2000 3 45,000 18 Air 200 14 1,440 Component tests<br />

Russia AA-X-12a DR 1995– 3.5 70,000 65 Air 142 8 390 Flight tests<br />

Israil Gabriel IVb LFIRR 1995– 2.5 0 320 Rail 190 17.6 2,100 Component tests<br />

People’s Republic <strong>of</strong> China Hsiung Feng IIIb LFIRR 1995– 2.5 0 320 Rail 190 17.6 2,100 Component tests<br />

United Kingdom FMRAAMb VFDR 1995–1999 3 70,000 100 Air 144 7 400 Development<br />

France Vestab LFIRR 1996–2003 4 0 186 Air 200 14 1,848 Flight tests<br />

France Rascalb LFIRR 1996– 2–4 0–65,000 —— —— —— 8–14 —— Component tests<br />

France ASMP-Ab LFIRR 1998– 4 0 270 Air 200 14 1,848 To be Operational<br />

Germany ARMIGERb DRIRR 1996– 3 30,000 100 Air 156–173 8 500 Component tests<br />

Russia SS-N-26/Yakhonta LFIRR 1998– 2.5 0–50,000 60–260 Rail 350 27.5 5,500 Flight tests<br />

India/Russia PJ-10/Brahmosb LFIRR 1998– 2.8 0–50,000 162 Rail 354 28 8,000 To be Operational<br />

United Kingdom BVRAAM/Meteora VFDR 1999– 4 50,000 54 Air 144 7 400 Component tests<br />

U.S. Navy SFRJ Tech Pgm SFRJ 1999– 5–6 0–80,000 1,000 Rail/VLS 168–256 9 2,200–3,700 Component tests<br />

2000–2003 France MICA/RJ f<br />

LFIRR 2000– 3 50,000 50 Rail 144 7.1 370 Component tests<br />

United Kingdom/Sweden S225XR VFDR 2000– 3 50,000 60 Rail 144 7.5 400 Component tests<br />

People’s Republic <strong>of</strong> China Ying-Ji 12b LFIRR 2000– 3 20,000 18 Air 200 13.8 1,440 Component tests<br />

U.S. Navy SSSTa VFDR 2000– 2.5+ 0 45–60 Rail 192–228 13.8 —— Flight tests<br />

U.S. Navy GSSCMb LFIRR 2002– 4.5 80,000 450 Rail/VLS 168 21 2,280 Component tests<br />

U.S. Navy HSADb VFDR 2002– 4.5 80,000 100 Air —— 10 —— Component tests<br />

c Anti-navire supersonique (ANS). d Boron solid fuel ramjet (BSFRJ). e Boron engine component integration, test and evaluation (BECITE). f Missile intermediat de combat aerien/ramjet<br />

a b System discussed and shown. System discussed.<br />

(MICA/RJ).


Table 5 Worldwide scramjet evolution 1955–1990<br />

Cruise Powered Total Total<br />

Engine Cruise altitude, range, length, Diameter, weight, State <strong>of</strong><br />

Era Country/service Engine/vehicle type Dates, year Mach no. ft n mile Launcher in. in. lbm development<br />

1955–1975 U.S. Navy External burnb ERJ 1957–1962 5–7 —— —— —— —— —— —— Combustion tests<br />

Russia Chetinkov research ERJ 1957–1960 5–7 —— —— —— —— —— —— Component tests<br />

U.S. Air Force Marquardt SJ DMSJ 1960–1970 3–5 —— —— —— 88 10 × 15 —— Cooled engine tests<br />

U.S. Air Force GASL SJa SJ 1961–1968 3–12 —— —— —— 40 31 in2 —— Cooled engine tests<br />

U.S. Navy SCRAMb LFSJ 1962–1977 7.5 100,000 350 Rail 288 26.2 5,470 Free-jet test<br />

U.S. Air Force IFTVc H2/SJ 1965–1967 5–6 56,000 —— —— —— —— —— Component tests<br />

U.S. Air Force–NASA HREa H2/SJ 1966–1974 4–7 —— —— —— 87 18 —— Flowpath tests<br />

NASA AIMb H2/SJ 1970–1984 4–7 —— —— —— 87 18 —— Cooled engine tests<br />

France ESOPEb DMSJ 1973–1974 5–7 —— —— —— 87 18 —— Component tests<br />

1975–1990 U.S. Navy WADM/HyWADMb DCR 1977–1986 4–6 80,000–100,000 500–900 VLS 256 21 3,750 Component tests<br />

Russia Various research SJ/DCR 1980–1991 5–7 80,000–100,000 —— —— —— —— —— Combustion tests<br />

NASA NASPb MCSJ 1986–1994 0–26 0–orbit Orbital Runway —— —— 500,000 Free-jet test (M7)<br />

Germany Sänger IIb ATRJ 1988–1994 4 0–orbit Orbital Runway 3976 550 800,000 Concept vehicle<br />

c IFTV incremental flight test vehicle.<br />

discussed.<br />

a System discussed and shown. b System<br />

FRY 39<br />

Fig. 9 France’s first ramjet-propelled flight tested airplane (1933–<br />

1951).<br />

Fig. 10 German Sänger I Hypersonic Concept Vehicle (1936–1945). 34<br />

as an air-launched vehicle from another aircraft. Refined versions<br />

<strong>of</strong> this aircraft followed over the years with the subsequent development<br />

<strong>of</strong> the Griffin II. Throughout this chronology, the dates in the<br />

figure captions are intended to reflect a total time interval including<br />

development, testing, and operation if applicable.<br />

In the late 1930s, Eugen Sänger, one <strong>of</strong> Germany’s top theoreticians<br />

on hypersonic dynamics and ramjets, and his wife, mathematician<br />

Irene Bredt, had begun developing a suborbital rocket<br />

bomber, Sänger I (Fig. 10) or RaBo, (sometimes called the Antipodal<br />

Bomber) that would be capable <strong>of</strong> attacking targets at intercontinental<br />

ranges. 34 Highly advanced concepts, including swept,<br />

wedge-shaped supersonic airfoils, a flat, heat-dissipating fuselage<br />

undersurface that anticipated the space shuttle’s design by 30 years,<br />

and rocket engines <strong>of</strong> extraordinary thrust were incorporated in the<br />

RaBo concept that would have taken many years to develop. In 1944,<br />

Sänger and Bredt were moved to an isolated laboratory complex in<br />

the mountains near L<strong>of</strong>er, Austria, where a number <strong>of</strong> advanced<br />

research projects were underway. Most sources indicate that nothing<br />

much came <strong>of</strong> the Amerika Bomber project at L<strong>of</strong>er, but this<br />

is clearly not the case. The Russians recovered copies <strong>of</strong> Sänger’s<br />

RaBo reports and were so fascinated with the concept that they<br />

dedicated effort to designing an updated RaBo (Antipodal Bomber)<br />

equipped with huge ramjet engines for boost and cruise propulsion.<br />

The RaBo influenced Soviet manned and unmanned rocket work<br />

for years after the war. It influenced U.S. work, too, leading directly<br />

to the Walter Dornberger-sponsored Bell bomber missile (BoMi)<br />

project <strong>of</strong> the early 1950s, and ultimately the U.S. Air Force/Boeing<br />

X-20 Dyna Soar hypersonic glider program that laid the technical<br />

groundwork for the space shuttle. The Sänger work to achieve Earth<br />

orbit recognized the need for airbreathing propulsion and suggested<br />

a preference for a SSTO vehicle.<br />

Both the Sänger–Bredt RaBo <strong>of</strong> 1945 and the postwar Soviet<br />

derivative used a long rocket sled to propel the vehicle to its take<strong>of</strong>f<br />

speed <strong>of</strong> several hundred miles per hour. After launch, an onboard<br />

engine <strong>of</strong> some 200,000-lb thrust would propel the craft into a ballistic<br />

trajectory that peaked at altitudes <strong>of</strong> several hundred miles. The<br />

German version was intended to be able to reach friendly territory<br />

after making its strike, or possibly ditch near a U-boat. The Soviet<br />

version had ramjets to provide ascent boost and possibly returnto-base<br />

cruise capability after velocity decayed into the supersonic<br />

range. After the RaBo arced to altitudes <strong>of</strong> several hundred miles


40 FRY<br />

Table 6 Worldwide scramjet evolution 1990–today<br />

Cruise Powered Total Total<br />

Engine Cruise altitude, range, length, Diameter, weight, State <strong>of</strong><br />

Era Country/service Engine/vehicle type Dates, year Mach no. ft n mile Launcher in. in. lbm development<br />

1990–2003 United Kingdom HOTOL c<br />

SJ 1990–1994 2–8 —— —— —— —— —— —— Combustion tests<br />

Japan PATRES/ATREXb TRBCC 1990– 0–12 100,000 —— —— 87 30 —— Component tests<br />

Japan NAL-KPL researchb SJ 1991– 4–12 50,000-100,000 —— —— 83 8 × 10 —— Component tests<br />

Russia Kholod a<br />

DCR 1991–1998 3.5–5.4 50,000–115,000 —— SA-5 36 24 —— Flight tests<br />

Russia/France Kholod a<br />

DCR 1991–1995 3.5–5.4 50,000–115,000 —— SA-5 36 24 —— Flight tests<br />

Russia/United States Kholod a<br />

DCR 1994–1998 3.5–7 50,000–115,000 —— SA-5 36 24 —— Flight tests<br />

France CHAMIOSb SJ 1992–2000 6.5 —— —— —— —— 8 × 10 —— Component tests<br />

France Monomat DMSJ 1992–2000 4–7.5 —— —— —— —— 4 × 4 —— Component tests<br />

France PREPAa DMSJ 1992–1999 2–12 0–130,000 Orbital Ground 2560 Waverider 1 × 106 Component tests<br />

Russia ORYOL/MIKAKS SJ 1993– 0–12 0–130,000 Orbital Ground —— —— —— Component tests<br />

France/Russia WRRb DMSJ 1993– 3–12 0–130,000 —— Ground —— Waverider 60,000 Component tests<br />

Russia GELA Phase IIa RJ/SJ 1995– 3–5+ 295,000 —— Tu-22M —— —— —— Flight tests<br />

Russia AJAXb SJ 1995– 0–12 0–130,000 —— —— —— —— —— Concept<br />

U.S. Air Force HyTecha SJ 1995– 7–10 50,000–130,000 —— —— 87 9 × 12 —— Component tests<br />

United States GTX d<br />

RBCC e<br />

1995– 0–14 50,000–130,000 —— —— —— —— —— Component tests<br />

U.S. Navy Counterforce DCR 1995– 4–8 80,000–100,000 —— Air/VLS 256 21 3,750 Component tests<br />

NASA X-43A/Hyper-Xa H2/SJ 1995– 7–10 100,000 200 Pegasus 148 60(span) 3,000 Flight tests<br />

France/Germany JAPHARa DMSJ 1997–2002 5–7.6 80,000 —— —— 90 4 × 4 —— Component tests<br />

United States ARRMDb DCR 1997–2001 3–8 80,000 450–800 Rail/Air 168–256 21 2,200–3,770 Component tests<br />

Russia IGLAa SJ 1999– 5–14 82,000–164,000 —— SS-25 197 —— —— Flight tests<br />

NASA X-43Cb DMSJ 1999– 5–7 100,000 —— Pegasus —— 10.5 wide —— Component tests<br />

Unites States IHPTETb ATR 1999– 0–5 0–90,000 —— —— —— 15–40 —— Component tests<br />

United States RTAb TBCC 1999– 0–5 0–90,000 —— —— —— 15–40 —— Component tests<br />

France Prometheeb DMSJ 1999–2002 2–8 0–130,000 —— —— 238 —— 3,400 Component tests<br />

India AVATAR-Mb SJ 1999– 0–14 0–orbit Orbital Ground —— —— 18–25 ton Combustion tests<br />

United Kingdom HOTOL Phase II SJ 2000– 2–8 —— —— —— —— —— —— Component tests<br />

France PIAFb DMSJ 2000– 2–8 0–110,000 —— —— 53 8 × 2 —— Component tests<br />

United States MARIAH MHD/SJ 2001– 15 —— —— —— —— —— —— Combustion tests<br />

Australia HyShotb SJ 2001–2002 7.6 75,000–120,000 200 Terrior Orion 55 14 —— Flight tests<br />

United States Gun launch technology SJ 2001– —— —— —— Ground —— —— —— Flight tests<br />

United States ISTARb RBCC e<br />

2002–2003 2.4–7 0-orbit Orbital Ground 400 Waverider 20,000 Component tests<br />

United States X-43Bb RB/TBCC 2002–2003 0–10 100,000 200 Air 500 Waverider 24,000 Component tests<br />

Russia Mig-31 HFL<br />

Waverider<br />

b SJ/DCR 2002– 2–10 50,000–130,000 —— Mig–31 —— —— —— Planned flight tests<br />

United States HyFlya DCR 2002– 3–6.5 85,000–95000 600 F-4 225 19 2,360 Flight tests planned<br />

United States SEDa SJ 2003– 4.5–7 80,000 —— —— —— 9 wide —— Planned flight tests<br />

France LEAa SJ/DCR 2003–2012 4–8 80,000 —— Air —— —— —— Flight tests planned<br />

United States RCCFD b<br />

TBCC 2003– 0.7–7 0–orbit Orbital Ground 400 20,000 Flight tests planned<br />

c Horizontal tak<strong>of</strong>f and landing (HOTOL). d NASA Glenn Rc hydrogen fueled/cooled (GTX). e Reference vehicle designation (RBCC).<br />

discussed.<br />

a System discussed and shown. b System


at speeds <strong>of</strong> over 10,000 mph, the spacecraft would descend and<br />

encounter the upper atmosphere. The pilot would execute a high-g<br />

pullout as the craft skipped <strong>of</strong>f the atmosphere like a stone skipping<br />

on water. This cycle would be repeated several times as the vehicle<br />

slowed and descended on its global path. When it neared its target,<br />

the spacecraft would release a large conventional-explosive bomb<br />

that would enter the atmosphere at meteor speeds and strike with<br />

tremendous force. No reference source on the Sänger–Bredt project<br />

indicates that any RaBo hardware was built at L<strong>of</strong>er. A 1947 U.S.<br />

technical intelligence manual on the L<strong>of</strong>er base, however, contains<br />

an image <strong>of</strong> an incomplete nose and forward fuselage <strong>of</strong> a very<br />

unconventional aircraft that was never flown.<br />

During and shortly after WWII, the United State developed<br />

a number <strong>of</strong> airbreathing-propelled drones. One example is the<br />

1942 McDonald-developed Katydid KDH-1 pulsejet-powered target<br />

drone, whose design heritage appears to follow from the German<br />

V-1. Unknown quantities <strong>of</strong> drones were produced.<br />

In the United States, the Marquardt Company, founded by R.<br />

Marquardt, began developing ramjet engines in 1944. Shortly after<br />

WWII, the U.S. military services developed and tested subsonic<br />

ramjet engines for a series <strong>of</strong> experimental applications. The propulsion<br />

goal in these early days <strong>of</strong> turbojet propulsion was to achieve<br />

low-cost airbreathing engines. Flight testing <strong>of</strong> these subsonic ramjets<br />

required only aircraft assist to reach ramjet flight speed. 35 Early<br />

interests in ramjet propulsion for high-speed subsonic aircraft applications<br />

led to two piloted experimental flight demonstrators. The<br />

first <strong>of</strong> these, conducted in 1946, employed a North American P-<br />

51 Mustang as the carrier aircraft with its dual-wingtip-mounted<br />

Marquardt subsonic ramjets. The second, flown in 1948, used the<br />

higher speed turbojet-powered Lockheed F-80 Shooting Star fighter.<br />

The ramjets were typically started fuel rich to achieve maximum<br />

thrust, accounting for the extended exhaust flames that diminished<br />

as flight speed increased. The F-80 briskly accelerated on ramjet<br />

power, with the turbojet set at idle. Larger 48-in.-diam experimental<br />

subsonic ramjets were subsequently developed and flown experimentally<br />

on the Douglas B-26 and demonstrated the potential for<br />

engine scale-up and added thrust. Subsonic ramjets were applied to<br />

production target drones and experimental missiles during this period.<br />

A leading example was the U.S. Navy Gorgon IV test vehicle<br />

first flown in 1947. Both this prototype missile and the derivative<br />

KDM-1 Plover target drone were powered by a single 20-in. ramjet<br />

burning gasoline and used parachute recovery permitting reuse <strong>of</strong><br />

the vehicle.<br />

The U.S. Navy Cobra was the first successful demonstration <strong>of</strong> a<br />

ramjet in supersonic flight, occurring in 1945. Cobra is a 6-in.-diam,<br />

normal shock inlet, tandem-boosted, liquid propylene oxide-fueled<br />

ramjet flight test vehicle that cruised at Mach 2 at 20,000-ft altitude.<br />

Its purpose was to demonstrate that a ramjet could produce the<br />

requisite thrust to cruise at supersonic speeds.<br />

The U.S. Navy Bureau <strong>of</strong> Aeronautics developed an entire family<br />

<strong>of</strong> unmanned guided drone missiles under the name <strong>of</strong> Gordon between<br />

1943 and 1953 for a variety <strong>of</strong> tasks, including ground attack<br />

and interception <strong>of</strong> bomber formations. In the mid-1940s, piston<br />

engines yielded insufficient performance and new ramjet and turbojet<br />

propulsion were beginning to show promise. Development <strong>of</strong><br />

a ramjet-powered Gordon IV began in 1945 using the Marquardt<br />

XRJ30-MA ramjet engine, a derivative <strong>of</strong> which supported the Bomarc<br />

missile development a few years later. The swept-wing KDM-1<br />

with an underslung ramjet first flew in 1947. There were 12 vehicles<br />

were produced and operational until 1949, when the hardware<br />

was used to develop a later version called Plover. The program was<br />

canceled in 1953 as advanced missile technology was developed.<br />

Considerably less sophisticated normal shock inlets were used<br />

in the early history <strong>of</strong> ramjet engines. These were the real flying<br />

stovepipes with no moving parts and advertised as the simplest<br />

propulsion system known to man. The focus shifted from the ultrasimplicity<br />

<strong>of</strong> the normal shock ramjet in the mid-1950s to the<br />

supersonic inlet ramjet engine as employed in the Bomarc.<br />

These subsonic ramjet flight test successes led to subsequent supersonic<br />

ramjet system developments. To do so, with only subsonic<br />

aircraft being available, required the use <strong>of</strong> rocket boosters. The op-<br />

FRY 41<br />

Fig. 11 U.S. NAVAHO operational missile (1946–1958).<br />

erational Bomarc and Talos interceptor missiles were derived from<br />

this advancement in propulsion technology. Marquardt developed<br />

the engines for the Bomarc missile, producing over 1600 units in<br />

various versions. The ramjet-powered Navaho and Talos missiles<br />

were developed in this same era. Escher and Foreman 36 assembled<br />

a recent excellent review <strong>of</strong> these missile systems and recount useful<br />

technical and program information.<br />

The Navaho supersonic cruise missile (Fig. 11), developed in the<br />

1950s for intercontinental ranges, was massive in size. At 48-in.<br />

diameter, 90 ft long and weighing in at 120,000 lbm, it is the largest<br />

ramjet engine developed in the United States. The missile used two<br />

engines mounted side-by-side, with each connected to an airflow<br />

duct leading from each <strong>of</strong> the two fuselage-side-mounted half-cone<br />

equipped inlets. The engines burned JP-4 or JP-5. The development<br />

program was canceled in 1957 due to its high cost together with<br />

improvements in surface-to-air missile (SAM) technology and the<br />

successful development <strong>of</strong> intercontinental ballistic missiles over<br />

supersonic cruise missiles.<br />

Russia developed the Burya (storm) missile in parallel with the<br />

U.S. Navaho. Designed to be an intercontinental strategic missile,<br />

the Burya employed an LFRJ cycle and two tandem boosters that<br />

used kerosene/nitric acid fuel. Five successful flight tests <strong>of</strong> the<br />

Mach 3 missile reportedly occurred. The development was canceled<br />

in 1958, shortly after development <strong>of</strong> the Navaho was stopped.<br />

The Bomarc and Talos antiaircraft interceptor systems were fully<br />

developed and operationally deployed by the U.S. Air Force and<br />

U.S. Navy, respectively. These rocket-boosted hydrocarbon-fueled<br />

ramjet-powered vehicles and derivatives routinely achieved speeds<br />

<strong>of</strong> Mach 2–3+ and altitudes from sea level to 40,000–70,000 ft and<br />

higher. In doing so, they recorded high degrees <strong>of</strong> mission reliability<br />

and operational effectiveness. The Bomarc supersonic interceptor<br />

missile (Fig. 12), deployed in the early 1960s by the U.S. Air Force<br />

and the Royal Canadian Air Force, employed two variations <strong>of</strong> a<br />

fixed-geometry ramjet engine. The A-series missile employed an<br />

axisymmetric cone-configured mixed compression inlet and annular<br />

air-cooled combustor that burned 80-octane gasoline. The B-series<br />

missile was fitted with an isentropic-configured fixed-inlet spike and<br />

burned JP-4 fuel. Launched vertically, the Bomarc interceptor rolled<br />

to its flyout azimuth, pitched over and accelerated to flight speeds <strong>of</strong><br />

Mach 2.5–2.7 and altitudes <strong>of</strong> 65,000–70,000 ft under ramjet power.<br />

The ramjet engine was later certified for Mach 3.4 service.<br />

In the early 1950s, after the need for a supersonic experimental<br />

research and flight demonstration vehicle was recognized, the<br />

United States initiated the X-7 program. This air-launched solidrocket-boosted<br />

supersonic vehicle initially used several Bomarc<br />

hydrocarbon-fueled conventional ramjet engines ranging in size<br />

from 28 to 36 in. in diameter and achieved flight conditions <strong>of</strong> Mach<br />

4.3 and 84,000 ft. Flown from 1952 to 1959, the X-7 was designed


42 FRY<br />

Fig. 12 U.S. BOMARC operational missile (1950–1972).<br />

Fig. 13 U.S. Talos/Vandal operational missile (1950–today).<br />

for recovery and reuse and was later adapted as a ground-launched<br />

supersonic target for the U.S. Army.<br />

The Talos supersonic interceptor missile (Fig. 13) evolved from<br />

the Bumblebee program under the direction <strong>of</strong> the JHU/APL initiated<br />

at the end <strong>of</strong> WWII. The Talos engine used a single, oncenterline<br />

ramjet engine with a double-cone configuration mixed<br />

compression inlet that fed to an air-cooled can combustor, which<br />

burned JP-5 fuel. Boosted to supersonic ramjet-takeover conditions<br />

by a separate solid propellant rocket, this ramjet-powered<br />

missile further accelerated and then cruised out supersonically to<br />

the aircraft-intercept point at high altitude. Its main limitations were<br />

in terminal intercept phase due to guidance inaccuracies and the axisymmetric<br />

inlet would unstart at high angles <strong>of</strong> attack followed by<br />

loss <strong>of</strong> control. The Talos was first fired in 1951 and introduced<br />

into the U.S. fleet in 1955. After the cancellation <strong>of</strong> the BQM-90<br />

program in 1973, the U.S. Navy had to look for other target missiles<br />

to simulate attacking antiship missiles. In 1975, it was decided to<br />

convert some obsolete RIM-8 Talos missiles to MQM-8G Vandal<br />

targets as a short-term solution to simulate the terminal phase <strong>of</strong> a<br />

missile attack. Some 2400 Talos units were built, with the U.S. Navy<br />

presently using a Talos variant (Vandal) as a low-altitude supersonic<br />

target. This is the second and final operational ramjet-powered system<br />

fielded by the United States to date and the only ramjet-powered<br />

system in limited flight service today.<br />

Refined versions <strong>of</strong> the Leduc-010, whose first flight was in 1949,<br />

followed over the years using air-turboramjet cycles to provide static<br />

thrust, with the most striking demonstration <strong>of</strong> ramjet supersonic<br />

flight speeds being achieved in the Griffon II from 1957 to 1961,<br />

which reached climb speeds <strong>of</strong> Mach 2.19 and Mach 2.1 at 50,000 ft<br />

(15.3 km) in 1959. Work on ramjets for aircraft applications generally<br />

stopped in the 1960s when the turbojet design was perfected<br />

and was shown to have lower fuel consumption than ramjets. Following<br />

the initial desires to employ ramjet technology in aircraft,<br />

subsequent development focused on advancing missile technology<br />

In 1955 the U.S. imitiated one <strong>of</strong> the first SFIRRs ever developed,<br />

ram air rocket engine, RARE. It employed a conical nose inlet<br />

and rocket-ramjet flowpath; the ramjet combustor used the chamber<br />

that housed the booster grain, but had a separate upstream chamber<br />

through which the air and solid fuel mixed and ignited. This dual<br />

in-line combustion chamber approach was used to solve potential<br />

boost-to-sustain transition problems. Magnesium and boron-loaded<br />

fuels were examined. Three flight tests <strong>of</strong> the RARE vehicle were<br />

successfully conducted at Mach 2.3 between 1959–1960.<br />

In the 1955–1965 time frame, France developed, flight tested, and<br />

made operational three LFRJ missiles: the VEGA, the SE-4400, and<br />

the target vehicle, CT-41.<br />

The U.S. Typhon development began in 1957. The engine employed<br />

an LFRJ design that used a Talos tandem solid propellant<br />

booster. The Typhon missile was much smaller than its Talos predecessor,<br />

yet still capable <strong>of</strong> flying almost twice the range. In addition<br />

to the reduced sustainer weight (1800 vs 3360 lbm for the Talos) the<br />

conical-inlet can combustor ramjet propulsion was more efficient<br />

and its subsystems and structure were more compact. It was successfully<br />

flight tested nine times from 1961 to 1963. Unfortunately,<br />

the Typhon was not introduced into the fleet because it outperformed<br />

the capabilities <strong>of</strong> the radar, guidance, control, and battle space coverage<br />

<strong>of</strong> the time. Despite its success, the program was canceled in<br />

1965. However, the lessons learned and technology developed were<br />

to become the cornerstones <strong>of</strong> the Aegis weapon system 10 years<br />

later.<br />

The U.S. Creative Research on Weapons (CROW) was developed<br />

with the goal <strong>of</strong> demonstrating SFIRR feasibility for delivering an<br />

air-launched payload to a desired location. 22 The axisymmetric IRR<br />

configuration used an SFRJ sustainer and an integral booster package<br />

contained within the ramjet combustor. The design employed<br />

ejectables including a bulkhead between the sustainer and booster<br />

grains and a rocket nozzle plug centerbody, which were expelled<br />

by ram air at transition. Six flight tests were successfully conducted<br />

demonstrating operational potential. The CROW concept, although<br />

briefly considered for use as an air-to-air missile and a high-speed<br />

target, never became operational.<br />

The Bloodhound (Fig. 14) and the later Sea Dart (Fig. 15) were<br />

developed and used by the Royal Navy and employed in the role <strong>of</strong><br />

long-range air defense. The Bloodhound is most easily recognized<br />

by its pr<strong>of</strong>usion <strong>of</strong> control surfaces and stabilizers; it has one fin<br />

on each <strong>of</strong> four boosters, two pivoting mainplanes, and two fixed<br />

horizontal stabilizers mounted in line with the wings. Two solid<br />

propellant boosters and an LFRJ sustainer propel the missile. The<br />

Bloodhound maneuvers with its mainplanes employing a twist-andsteer<br />

technique. Initially operational in 1964, almost 800 missiles<br />

were produced until 1991, when it was removed from service.<br />

<strong>Ramjet</strong> Development: 1960–1980<br />

A major milestone <strong>of</strong> this 20-year period was the conception and<br />

demonstration <strong>of</strong> IRR technology for missile applications. Table 3<br />

summarizes ramjet evolution in the era from 1960 to 1980 and provides<br />

originating country, engine/vehicle names, development dates,<br />

performance, physical characteristics, and state <strong>of</strong> development. The


Fig. 14 U.K. Bloodhound operational missile (1957–1991).<br />

Fig. 15 U.K. Sea Dart operational SAM (1960–1975).<br />

general pace <strong>of</strong> ramjet development accelerated during this era, with<br />

the People’s Republic <strong>of</strong> China being added to the countries engaged<br />

in activities. As one witnesses the dramatic growth due to<br />

competition between maturing siblings, thus it is with the growth in<br />

international ramjet technology.<br />

France developed the experimental LFRJ Stataletex missile from<br />

1960 to 1970, which achieved Mach 5 at 82,000 ft in flight tests at<br />

the end <strong>of</strong> the 1960s.<br />

To field a fully responsive target to validate the surface-to-air<br />

Hawk interceptor missile system, the U.S. Army sponsored development<br />

<strong>of</strong> the specially tailored MQM-42 Redhead Roadrunner tandem<br />

rocket-boosted ramjet-powered target vehicle in the mid-1960s.<br />

The old normal shock ramjet reappeared again in this system, which<br />

was designed to simulate Mach 0.9–1.5 hostile aircraft penetrating<br />

defenses at medium to low altitudes. Like the KDM-1 and X-7,<br />

this vehicle was designed for parachute recovery. The cost-savings<br />

pay<strong>of</strong>f <strong>of</strong> reusability was demonstrated with over 90 flights accomplished<br />

with some 50 sets <strong>of</strong> flight hardware.<br />

The United Kingdom began development <strong>of</strong> the Sea Dart (Fig. 15)<br />

as a lightweight area defense replacement for the Sea Slug in 1960.<br />

The missile employs a solid propellant booster and an LFRJ sustainer.<br />

Flight testing began in 1965, and if was operational as early<br />

as 1967. The Sea Dart accounted for seven kills in the Falkland<br />

conflict and is credited with downing an Iraqi missile headed for<br />

a U.S. cruiser in the Gulf War <strong>of</strong> 1991. The Argentineans reportedly<br />

ordered 100 missiles in 1986. Currently, the Sea Dart is being<br />

removed from Royal Navy service.<br />

One <strong>of</strong> the earliest attempts in the United States to integrate the<br />

ramjet with a liquid propellant booster was the Hyperjet concept,<br />

flight tested at speeds over Mach 5 in the early 1960s using hydrogen<br />

peroxide and kerosene. Following rocket boost, the engine<br />

would transition to and accelerate in the ramjet mode. Although<br />

the Hyperjet’s liquid propellant boost mode was repeatedly demonstrated<br />

in flight, logistics and field handling subsequently encouraged<br />

development <strong>of</strong> alternative designs.<br />

In the late 1960s, the United States successfully developed and<br />

flight tested a small, high-altitude Mach 3+ reconnaissance vehicle<br />

designated the D-21 (Fig. 16). It was powered by a closely<br />

vehicle-integrated, high-speed version <strong>of</strong> the Bomarc RJ-43-MA<br />

FRY 43<br />

Fig. 16 U.S. D-21 operational reconnaissance vehicle (1963–1980).<br />

Fig. 17 Russian SA-4 operational SAM (1964–today).<br />

Fig. 18 Russian SA-6 operational SAM (1965–today).<br />

JP-7-fueled LFRJ engine and was designed for carriage and launch<br />

from the SR-71 Blackbird at supersonic flight conditions. The engine<br />

design capitalized on the Bomarc production design. By 1972, it had<br />

flown operationally, achieving altitudes <strong>of</strong> 92,000 ft and ranges over<br />

3000 miles. The D-21 was designed for supersonic launch from an<br />

SR-71 Blackbird aircraft and programmed to fly a racetrack-type reconnaissance<br />

course over large land masses at altitudes in excess <strong>of</strong><br />

80,000 feet and speed over Mach 3. On return to the launch zone, the<br />

reconnaissance package was to be jettisoned for midair parachute<br />

recovery. The operational design was eventually launched from the<br />

B-52.<br />

<strong>Ramjet</strong>-powered SAM development began in the early 1960s in<br />

Russia with the SA-4 Ganef, a solid propellant boosted LFRJ cycle<br />

that became operational by 1967 (Fig. 17). At least four variants <strong>of</strong><br />

the medium to high altitude missile were produced between 1967<br />

and 1973. The missile uses four solid propellant boosters mounted<br />

externally on the kerosene-fueled sustainer body.<br />

Development <strong>of</strong> the Russian SA-6 Gainful second generation<br />

SAM began in the mid-1960s and was operational by 1970 (Fig. 18).


44 FRY<br />

It uses a DR IRR cycle and is commonly attributed with starting<br />

the ramjet race. The October 1973 Yom Kippur War convincingly<br />

demonstrated the effectiveness <strong>of</strong> the IRR in tactical missiles. With<br />

employment <strong>of</strong> a DR cycle, the Russian SA-6 SAMs inflicted sharp<br />

losses on the Israeli fighters in 1973 and played prominently in the<br />

1999 air campaign in Yugoslavia.<br />

Before 1965, ramjet-powered missile systems, such as the U.S.<br />

Air Force Navaho and Bomarc, the U.S. Navy Talos, and the Russian<br />

SA-4 were externally boosted by a separable rocket system. This led<br />

to relatively large missiles, which were limited to launch from the<br />

ground or a ship. Launch from an aircraft or greater maneuvering<br />

capability, which require a smaller missile, became feasible with the<br />

development <strong>of</strong> the integral rocket–ramjet concept. In these more<br />

recent ramjets, the rocket booster was integrated within the ramjet<br />

combustor, thereby saving considerable volume (Fig. 19). 37<br />

As with the axisymmetric normal shock tandem-boosted ramjet,<br />

the flow-turning inlet integral rocket ramjet has undergone cyclical<br />

periods <strong>of</strong> interest, from the supersonic chemical propulsion<br />

low-altitude supersonic ramjet missile (SCP/LASRM) in the mid-<br />

1960s to advanced low-volume ramjet (ALVRJ) in the early 1970s<br />

to advanced strategic air-launched missile–propulsion test vehicle<br />

(ASALM–PTV) in the late 1970s to the SLAT in the early<br />

1990s. <strong>Ramjet</strong>-powered air-to-surface missile exploratory development<br />

and flight testing began in the mid-1960s and extended<br />

through the 1980s. Solid and liquid IRR propulsion were selected<br />

for such applications. Beginning in 1964, the U.S. Air Force sponsored<br />

SCP/LASRM, an air-launched IRR supersonic missile development<br />

program. LASRM, the first low-volume IRR design, was<br />

flight demonstrated in the mid-1960s. The flight vehicle employed<br />

four aft-mounted, two-dimensional, inward turning inlets, was fueled<br />

with a heavy hydrocarbon, Shelldyne, and was designed to<br />

operate at Mach 2.5 sea level. The LASRM physical configuration<br />

is compared with contemporary systems in Fig. 19. Whereas flight<br />

tests were successful, a small operational envelope was experienced<br />

due to inlet unstart conditions.<br />

France developed the SE-X-422 demonstrator (Fig. 20) from 1965<br />

to 1967 for a long-range cruise missile application. The design employed<br />

an LRFJ with a single inlet located aft (Fig. 20). Three successful<br />

flight tests were conducted in 1967 and achieved Mach 2 at<br />

30,000 ft.<br />

The Navy ALVRJ vehicle (Fig. 21) was an LFIRR configuration<br />

<strong>of</strong> the late 1960s. The major program goal was to demonstrate controlled<br />

free flight at Mach 2.5–3 to ranges <strong>of</strong> 25–100 n mile with sufficient<br />

margin to provide for terminally effective tactical payloads. 22<br />

It was similar in size and configuration to LASRM, except the four<br />

aft-mounted side inlets were outward turning, which provided better<br />

angle-<strong>of</strong>-attack operation, thereby permitting operation from sea<br />

Fig. 19 Early ramjet systems history (1945–1980).<br />

Fig. 20 French SE-X-422 flight-tested missile (1965–1967).<br />

Fig. 21 U.S. ALVRJ flight-tested missile (1968–1975).<br />

level to intermediate altitudes. The JP-5-fueled ramjet employed<br />

a solid propellant carboxyl-terminated polybutadiene/ammonium<br />

perchlorate (CTPB/AP) integral booster, a DC93-104 insulation system,<br />

and an ejectable nozzle. Seven flight tests were flown, most<br />

between 1975 and 1979. Initial flight tests used flight weight heatsink<br />

hardware, with later testing involving the DC93-104 insulation<br />

system. All goals, objectives, and specified data points were substantially<br />

achieved with the major contribution being the first successful<br />

demonstration <strong>of</strong> the in-flight ramjet takeover for an IRR design. A<br />

number <strong>of</strong> other advanced development efforts to support a technology<br />

base for a supersonic tactical missile (STM) were conducted<br />

in parallel with the ALVRJ effort. Successful programs were conducted<br />

in the areas <strong>of</strong> terminal guidance, midcourse guidance, and<br />

warheads. Whereas the Navy approved an STM concept directed<br />

toward tactical land targets, Congress canceled subsequent development<br />

based upon a review <strong>of</strong> tactical needs and requirements.<br />

Russia began development <strong>of</strong> the SS-N-19 Shipwreck supersonic<br />

long-range antiship missile (Fig. 22) in the early 1970s. First thought<br />

to use turbojet propulsion, it was recently revealed 38 to be liquid fuel<br />

ramjet-powered. Made operational in 1981 as the first sublaunched


Fig. 22 Russian SS-N-19 operational missile (1972–today).<br />

antiship ramjet missile, the aptly named missile is in service today.<br />

It is deployed in both ships and submarines. The submarine Kursk,<br />

lost in the Barents Sea in August 2000, was reportedly conducting<br />

SS-N-19 live firings before it went down.<br />

The generic ordnance ramjet engine (GORJE) missile was developed<br />

through engine testing from 1972 to 1976 as an airbreathing<br />

propulsion system for the high-speed antiradiation missile (HARM).<br />

Such a ramjet design could provide higher average velocity with a<br />

lower peak velocity than a solid propellant rocket. This was viewed<br />

as a solution to relieving the then-current seeker dome aeroheating<br />

problem. An interesting parallel rocket ramjet configuration was devised,<br />

which located the RJ-4 fuel tank forward <strong>of</strong> the solid booster<br />

followed by the ramjet combustor positioned aft. A short L/D (∼1)<br />

annular ramjet combustor, which employed DC93-104 insulation,<br />

was used around a central blast tube for booster exhaust. The design<br />

employed four side-mounted axisymmetric inlets with the entrance<br />

just aft <strong>of</strong> the trailing edge <strong>of</strong> the midbody control wings. Aircraft<br />

interface requirements dictated this inlet configuration, which created<br />

inlet entrance flow distortion that was to plague the design<br />

performance. Because this design did not employ a traditional IRR<br />

configuration, neither combustor port covers nor inlet covers were<br />

necessary. Development <strong>of</strong> the LFIRR propulsion system was taken<br />

through semifreejet engine testing and was plagued with combustor<br />

oscillations and instability problems. Even though most <strong>of</strong> the<br />

expected pressure oscillations were accommodated by the inlet pressure<br />

recovery margin, some unanticipated higher frequency instabilities<br />

unstarted the inlets. The program provided impetus for development<br />

activities to understand causes and control methods <strong>of</strong><br />

combustion instabilities in short L/D LFRJ engines. Planned flight<br />

tests were never conducted due to lack <strong>of</strong> funds and the development<br />

was concluded in 1976.<br />

In the early 1970s, France pursued the exploratory development<br />

on Scorpion for a long range SAM mission. 39 This missile used a<br />

subsonic combustion LFRJ operating up to Mach 6 after a tandem<br />

boost to Mach 3 takeover. Successful ground testing was accomplished.<br />

Although this work did not proceed to flight demonstration,<br />

the analysis and experimental work demonstrated an important advance<br />

in high-speed ramjet technology.<br />

The Russians employed a first generation flight test vehicle GELA<br />

phase 1 testbed. 64−66 This phase dealt with Mach 3 ramjet missile<br />

propulsion systems, which were developed from 1973 to 1978 and<br />

ultimately used on the SS-N-22 and SA-6 upgrades. Limited information<br />

is available on these early flight-test activities.<br />

The potential <strong>of</strong> the IRR concept as a viable missile propulsion<br />

system attracted much attention, particularly with the advent <strong>of</strong> the<br />

Russian SA-6. The IRR concept enabled a compact air-launched missile<br />

configuration as shown in Fig. 19. Although most <strong>of</strong> these IRR<br />

systems operated at modest Mach number, the U.S. air-launched<br />

ASALM missile was tested to flight speeds approaching hypersonic<br />

Mach numbers. The U.S. Air Force sponsored the ASALM–PTV<br />

program and provided the most impressive flight demonstrations <strong>of</strong><br />

the capability <strong>of</strong> the air-launched IRR. 5,40 Begun in 1968, the first<br />

ground test was in 1975. This propulsion system used the IRR approach<br />

first tested in the LASRM also under U.S. Air Force sponsorship,<br />

but with significant design differences. This vehicle employed<br />

FRY 45<br />

Fig. 23 U.S. ASALM flight-test vehicle (1965–1980).<br />

a forward mounted “chin” inlet, which has the characteristics <strong>of</strong><br />

improved packaging and improved performance at wider angles <strong>of</strong><br />

attack (Fig. 23). 41 This permitted operation over a very wide flight<br />

envelope from Mach 2.5 sea level to Mach 4 at 80,000 ft. This inlet<br />

fed an ablatively cooled ramjet engine. ASALM tests demonstrated<br />

the capability <strong>of</strong> igniting ramjet fuel at −65 ◦ F and then achieving<br />

high combustion efficiency over a wide range <strong>of</strong> fuel/air ratios,<br />

as well as sustained operation at very low fuel/air ratios, which<br />

allowed for extended duration cruise. Whereas ramjet engine designs<br />

examined throughout these developments were challenged<br />

with operability concerns, combustion oscillations were relatively<br />

absent, unlike problems encountered in liquid and solid propellant<br />

rockets. ASALM was designed to fit the B-52 rotary launcher and<br />

provide significantly greater stand<strong>of</strong>f range to the rocket-powered<br />

short-range attack missile (SRAM) plus air-to-air capability against<br />

aircraft. Seven successful flight tests were conducted between October<br />

1979 and May 1980. Whereas MacDonald Douglas and Martin<br />

manufactured two different versions <strong>of</strong> the hardware for testing, the<br />

Martin/Marquardt version was flight tested. Unlike the similarly designed<br />

French ASMP, the ASALM did not see operational service,<br />

likely due to budget restrictions and the concurrent development <strong>of</strong><br />

the AGM-86 air-launched cruise missile (ALCM). Nevertheless, the<br />

successful ASALM flight tests showed the potential <strong>of</strong> a high-speed<br />

air-launched IRR missile that could be carried in a bomber aircraft<br />

and could possess long stand<strong>of</strong>f strike capability. 42 The marriage <strong>of</strong><br />

the high-speed capability <strong>of</strong> the ramjet and the compact air-launched<br />

missile was successfully demonstrated.<br />

The old flying stovepipe came back again in the late 1970s in<br />

the form <strong>of</strong> the U.S. Navy target vehicle Firebrand, used to simulate<br />

aircraft. Firebrand was designed as a parachute-recoverable<br />

ramjet-powered target suitable for ground and air launch. <strong>Propulsion</strong><br />

included tandem solid propellant boosters and two Marquardt<br />

LFRJ sustainers located in aft pods. The original plan was to build<br />

nine Firebrand flight-test vehicles and begin flight testing in 1983.<br />

The program encountered funding difficulties and the vehicle came<br />

in heavy for planned air launch from a C-130 aircraft. Whereas<br />

the Firebrand was ultimately canceled in 1983, it provided technology<br />

and engine hardware used in subsequent programs. One such<br />

example was the conceptual ramjet engine (CORE) engine, which<br />

used inner portions <strong>of</strong> the Firebrand, and was the first in its day to<br />

employ a conventional flameholder to be successfully tested at the<br />

aerodynamically severe conditions <strong>of</strong> Mach 2.5 sea level. Although<br />

Vandal targets had been in use since 1975, the U.S. Navy decided<br />

in 1983 to continue use <strong>of</strong> the Vandal as a target and formulated a<br />

new requirement for a dedicated antiship missile target. The latter<br />

eventually resulted in development <strong>of</strong> the AQM-127 SLAT missile.<br />

Russia began development <strong>of</strong> the SS-N-22 Sunburn (Fig. 24)<br />

in 1977. It was the first surface-to-surface missile (SSM) surfacelaunched<br />

(later air-launched) antiship missile with an initial operational<br />

capability (IOC) <strong>of</strong> 1982. This LFIRR engine design has had<br />

considerable influence on ramjet engine design and defenses against<br />

it capabilities. It is reported to be on the People’s Republic <strong>of</strong> China’s<br />

newly acquired Sovremenny-class guided-missile destroyers. The<br />

first Chinese test was reported 43 to be in 2001 and up to several hundred<br />

are potentially in their inventory today. This prompted Taiwan,<br />

Republic <strong>of</strong> China, to respond in 2001 with the development and


46 FRY<br />

flight testing <strong>of</strong> an LFRJ antiship supersonic missile, the Hsiung<br />

Feng III.<br />

In 1978, France began work on Rustique/MPSR-1, 44 which employs<br />

an unchoked flow ducted rocket (UFDR) engine and design<br />

concept oriented toward a mass-produced low-caliber missile<br />

with system advantages <strong>of</strong> storability, maintenance, and reliability<br />

characteristics comparable to those <strong>of</strong> solid propellant missiles.<br />

Rustique/MPSR-1 includes an air induction consisting <strong>of</strong> four twodimensional<br />

inlets, single igniter for a nozzleless booster, and unchoked<br />

gas generator. No choked throat exists between the gas generator<br />

and the ramjet combustion chamber. The gas generator solid<br />

propellant burning rate depends on the combustion chamber pressure<br />

and self-modulation <strong>of</strong> the mixture ratio established by correct<br />

selection <strong>of</strong> the solid propellant pressure exponent for the expected<br />

range <strong>of</strong> speed and altitude conditions. This propulsion concept was<br />

flight tested (five ground launches) through the MPSR-1 and again<br />

in 1993–1997 in MPSR-2 (two ground launches).<br />

In the late 1970s, in the People’s Republic <strong>of</strong> China, development<br />

began <strong>of</strong> the C-101 or CSS-C-5 Saples, a shore-based, supersonic,<br />

antiship missile similar in configuration to the British Bloodhound<br />

SAM. It can also be launched from air and ship platforms. The requirements<br />

<strong>of</strong> high speed and long range resulted in a large missile<br />

with two tandem solid propellant boosters and two ramjet sustainer<br />

engines. The ramjets are mounted on stubs extending from the aft<br />

fuselage sides. The boosters are nestled under the stubs, between<br />

the ramjets and the fuselage. A tritail vertical stabilizer group consisting<br />

<strong>of</strong> a single, rectangular vertical stabilizer mounted ahead <strong>of</strong><br />

abutterfly tail (two surfaces angled outward) provides control to the<br />

missile. A larger, longer ranged missile variant <strong>of</strong> the C-101, the<br />

C-301, was developed and fitted with four boosters.<br />

<strong>Ramjet</strong> Development: 1980–2000<br />

During this time frame, the intensity <strong>of</strong> international development<br />

activities increased in pace and number and moved convincingly into<br />

supersonic combustion. Table 4 summarizes ramjet evolution in the<br />

era from 1980 to 2000 and provides originating country, engine/<br />

vehicle names, development dates, performance, physical characteristics,<br />

and state <strong>of</strong> development. Whereas the countries engaged in<br />

ramjet developments did not expand during this era from those cited<br />

earlier, the international development activities did, which resulted<br />

in more operational systems. Wilson et al., 45 Dunfee and Hewitt, 46<br />

and Hewitt 47 provide recent reviews <strong>of</strong> worldwide developments<br />

in ramjet-powered missiles that have contributed to the following<br />

discussions <strong>of</strong> the developments during this era.<br />

France initiated development <strong>of</strong> ASMP in 1980 to satisfy a requirement<br />

for an air-launched nuclear stand<strong>of</strong>f missile. A competition<br />

between turbojet and ramjet propulsion preceded the start <strong>of</strong> this<br />

development. Flight-testing the French version <strong>of</strong> ASALM LFIRR<br />

Fig. 24 Russian SS-N-22 operational missile (1977–today).<br />

Fig. 25 U.S. SLAT flight-tested missile (1986–1992).<br />

technology occurred from 1980 to 1986. ASMP was subsequently<br />

deployed in 1986, was produced up to 1992, and is still in service<br />

today. Development <strong>of</strong> many variants has been pursued to varying<br />

degrees from 1991 to 2000 for air to air, air to surface (ASMP-<br />

C), and antiship anti-navire future/anti-navire nouvelle generation<br />

(ANF/ANNG) applications. Development <strong>of</strong> the air-to-air variant<br />

was shelved in the mid-1990s at the request <strong>of</strong> NATO partners. Development<br />

<strong>of</strong> the air-to-surface variant ended when development <strong>of</strong><br />

the turbojet-propelled SCALP was selected for this role in 1994.<br />

France and Germany pursued development <strong>of</strong> the antiship variant<br />

jointly from 1995 until completion <strong>of</strong> design studies in 1998 when<br />

Germany withdrew. The development reverted to its old name <strong>of</strong><br />

ANF and continued with plans for pro<strong>of</strong>-<strong>of</strong>-concept testing <strong>of</strong> the<br />

propulsion system on the testbed VESTA and deployment in 2005.<br />

Development was stopped in 2000 due to budget shortfalls, but the<br />

potential for program restart still remains. 48 Development <strong>of</strong> a replacement<br />

(ASMP-A) began in 1996 and is currently in progress<br />

with an expected IOC <strong>of</strong> 2006.<br />

The U.S. Navy began development <strong>of</strong> an advanced common intercept<br />

missile demonstrator (ACIMD) in 1981 as a long-range advanced<br />

antiair missile (AAM) replacement for the Phoenix. The<br />

LFIRR employed a two-dimensional midbody inlet and exhibited<br />

an ejectable solid booster nozzle. Development was canceled in<br />

1989 before flight tests due to a shortage <strong>of</strong> funding.<br />

The Firebrand flight conditions (Mach 2.5 sea level) were selected<br />

for the U.S. Navy AQM-127 SLAT development in the late<br />

1980s, with the IRR configuration ultimately being selected over<br />

the simple normal shock configuration. The technology base developed<br />

under the ASALM program was used in the development <strong>of</strong><br />

SLAT (Fig. 25) to provide aerial targets in support <strong>of</strong> test, evaluation,<br />

and training <strong>of</strong> shipboard defense systems. SLAT employed<br />

a supersonic single-duct chin inlet IRR configuration derived from<br />

the ASALM–PTV. Eight flight tests were conducted, but the program<br />

was canceled due to problems encountered in the flight tests,<br />

which were found to be unrelated to the propulsion. Development<br />

<strong>of</strong> this propulsion system is an example <strong>of</strong> successful transition <strong>of</strong><br />

technology between programs. The successful ramjet engine and its<br />

development framework can be used for future systems. After the<br />

cancellation <strong>of</strong> the earlier BQM-90 and BQM-111 Firebrand programs,<br />

SLAT was the Navy’s third attempt to develop a dedicated<br />

high-speed antiship missile threat simulator. A current effort in this<br />

area is the GQM-163 SSST.<br />

The United States began developing VFDR technology in the<br />

1980s. The VFDR missile (Fig. 26) concept was successfully developed<br />

from 1987 to 1997. The development, aimed at an advanced<br />

medium range air-to-air missile (AMRAAM) application, successfully<br />

took the design to the point <strong>of</strong> being ready for flight testing. The<br />

VFDR program included the following achievements: development


<strong>of</strong> an integrated propulsion system, flight-weight hardware and thermal/structural<br />

design, nozzles booster that demonstrated 103% <strong>of</strong><br />

design requirements in static firings for temperature cycling from<br />

−65 to 145 ◦ F, a flight-type gas generator and throttle valve assembly<br />

demonstrated in environmental tests, and sustainer combustion<br />

performance meeting or exceeding design requirements. More recently,<br />

an advanced design was tested in 2002 using a high-energy,<br />

reduced-smoke propellant in the two-inlet IRR configuration simulating<br />

conditions at Mach 3.25, 30,000 ft. The technology was not<br />

applied for various reasons until initiation <strong>of</strong> the U.S. Navy’s SSST<br />

development in 2000.<br />

In 1990, France initiated work on MARS, a ramjet-powered<br />

stealthy missile designed to be powered by an LFIRR and able to<br />

operate up to Mach 4. A model <strong>of</strong> MARS was displayed at the ILA<br />

2000 Air Show, with component tests being conducted on the engine<br />

technology, air intakes, and general missile aerodynamic configuration<br />

by 2002. 49 The last report suggested additional funding was<br />

necessary for development to continue.<br />

In the mid-1980s, India began indigenous development <strong>of</strong> a<br />

medium range SAM, the AKASH missile. Planned as a replacement<br />

for the SA-6 now in service, AKASH’s design is based in<br />

large part on the SA-6 DRIRR configuration. Development was repeatedly<br />

delayed with flight tests finally being initiated in late 1990.<br />

Flight testing continues today with tests reported in 2000, 2001, and<br />

2002. 50 Production to date appears to be on a very limited scale.<br />

India is currently assessing full production against the evaluation <strong>of</strong><br />

other SAM systems available on the world market for comparable<br />

requirements. 51<br />

In the early 1990s, in the People’s Republic <strong>of</strong> China, development<br />

began <strong>of</strong> the CSS-C-6 Sawhorse, a large, shore-based, antiship<br />

missile. It resembles a scaled-up C-101 CSS-C-5 Saples, but has a<br />

thicker fuselage. Although first seen in a Beijing display in November<br />

1988, the IOC is unknown. 52 Export versions are designated the<br />

C-301. The missile employs four, tandem solid propellant boosters<br />

located above and below each LFRJ sustainer engine. The two<br />

ramjets are mounted on narrow pylons extending from the sides <strong>of</strong><br />

the fuselage. A butterfly vertical tail stabilizes the missile, and each<br />

booster has its own single, angled stabilizer. Whereas it exhibits<br />

twice the range <strong>of</strong> the smaller C-101, the C-301 cruises at Mach 2<br />

with adjustable cruising altitudes. Although the missile is available<br />

for export, it is not believed to be in full operational Chinese service.<br />

Whereas the development <strong>of</strong> the Russian AS-17, Kh-31 Krypton<br />

(Fig. 27) began in the late 1970s, it was not until 1994 that it was<br />

deployed using LFIRR technology in an antiradiation missile configuration.<br />

It is in service today, with an upgraded air-launched version,<br />

the AS-17A/Kh-31P, and was revealed during a recent air show. 53<br />

Although the upgraded AS-17 is not yet available for export, such<br />

opportunities are being actively pursued.<br />

The Kh-31 made news in the mid-1990s, primarily due to its use<br />

as a missile target for the U.S. Navy. The Boeing Company imported<br />

airframes from Russia and modified them for use as MA-31 aerial<br />

targets. The first flight <strong>of</strong> an MA-31 occurred in mid-1996 with additional<br />

target vehicles under contract in 1999. However, Russian export<br />

restrictions imposed in 2001 made further deliveries uncertain.<br />

The People’s Republic <strong>of</strong> China and Russia are reportedly 54<br />

jointly developing the KR-1 antiradiation missile, based on the<br />

Kh-31P. China and Russia are cooperating on the development,<br />

with initial missile deliveries having occurred in 1997.<br />

In the early 1990s, South Africa began development <strong>of</strong> a DRIRR<br />

missile, LRAAM, with Israeli assistance. 46 Five preprototype flight<br />

tests had been conducted through 1995 on a four-inlet, nozzleless<br />

FRY 47<br />

Fig. 26 U.S. VFDR component-tested missile (1987–1997).<br />

Fig. 27 Russian AS-17 RJ operational missile (1994–today).<br />

booster, magnesium-based, fixed-flow (no throttling) DR configuration.<br />

Today, the design seems to have evolved toward a two-inlet,<br />

throttleable hydrocarbon fuel design, which is viewed as a Darter<br />

air-to-air missile (AAM) variant. An estimated IOC is 2005.<br />

The U.S. Navy initiated development <strong>of</strong> a Mach 4 low-drag<br />

ramjet/low-cost missile system (LDRJ/LCMS) Fasthawk in 1995<br />

with the expectation <strong>of</strong> flight testing in 1999. The system exhibited<br />

a forward concentric supersonic inlet and an LFRJ with a droppable<br />

solid propellant booster. Development demonstrated stable high efficiency<br />

in a short L/D combustor, a low-drag, roll control concept,<br />

and an efficient axisymmetric inlet design. Funding for the program<br />

was terminated in late 1998 despite showing good technical<br />

progress and demonstrating advances in structural design, control,<br />

and propulsion. Preparations for a flight-test demonstration were<br />

subsequently made in 1999, but the system has yet to be flown.<br />

In the mid-1990s Russia began development to incorporate ramjet<br />

propulsion into the AA-X-12 RVV-AE-PD, an improved longrange<br />

version <strong>of</strong> the medium range beyond-visual-range AAM the<br />

AA-12 Adder (Fig. 28). There were 5 firings reported in 1995 55 and<br />

10 ground tests reported in 1999, 56 with flight tests due to begin<br />

shortly thereafter. Reports 57 in 2001 suggested the baseline inlet<br />

configuration selected was a four-intake design; however, other options<br />

were still being considered, and projected fuel consumption<br />

was still higher than desired. A solid fuel DR engine is being used<br />

with automatic ram pressure controlled throttling. This seems to<br />

support the design advantages <strong>of</strong> an all-solids approach to this class<br />

<strong>of</strong> tactical missiles. This missile is the primary threat driving the<br />

U.K. BVRAAM missile requirement. An operational time frame <strong>of</strong><br />

2005 is projected.<br />

Israel began development <strong>of</strong> solid propellant surface-to-surface<br />

guided missiles in 1960. 58 When the Gabriel entered service a<br />

decade later, they were the first Western designed-to-purpose antiship<br />

missiles to become operational. Development <strong>of</strong> a ramjetpropelled<br />

Gabriel IV, which began in the early 1990s, is a larger missile<br />

with greater range than earlier variants. Development is believed<br />

to be in component testing, but has stalled since the mid-1990s. The<br />

propulsion includes an integral solid propellant booster and an airturboramjet<br />

sustainer engine. The Taiwanese SSM Hsiung Feng III,<br />

with an expected IOC <strong>of</strong> 2004, is a derivative <strong>of</strong> the Gabriel IV.<br />

In the early-to-mid-1990s after a comprehensive study <strong>of</strong> the<br />

new medium-range AAM (MRAAM) threat capabilities, the United


48 FRY<br />

Fig. 28 Russian AA-X-12 Adder RVV-AE-PD flight-tested missile (1995–today).<br />

Kingdom, Sweden, and Germany, acting both individually and in<br />

cooperation, were evaluating all-solids ramjet propulsion for a new<br />

future MRAAM FRAAM. 46 The requirements established for these<br />

new missiles were very similar, prompting their consolidation for<br />

many reasons. Four missile configurations were in competition for<br />

satisfying these requirements to provide a capability for the new Eur<strong>of</strong>ighter:<br />

the French MICA/RJ, the German A3M, the U.K./Sweden<br />

S225XR and the U.S. VFDR. This activity has evolved into the current<br />

development <strong>of</strong> the Meteor/BVRAAM. Germany is developing<br />

the propulsion system for the BVRAAM/Meteor, which employs a<br />

VFDR engine design likely derived from an earlier U.S./German<br />

cooperative development program conducted in 1989–1999.<br />

France has been developing IRR technology since 1996 in support<br />

<strong>of</strong> the requirements for the ANF supersonic antiship missile and will<br />

now be used to power the ASMP-A. A missile design capable <strong>of</strong><br />

operating at Mach 4–6 is expected to be available in 2006. Full-scale<br />

freejet tests at ramjet speeds <strong>of</strong> Mach 3 have been accomplished.<br />

France is using VESTA flight-test vehicles to demonstrate DR and<br />

LFIRR ramjet technology that could be applicable to a number <strong>of</strong><br />

programs. Three VESTA flight tests <strong>of</strong> the propulsion system were<br />

conducted in 2002–2003. The first ground launches <strong>of</strong> ASMP-A<br />

are scheduled for 2004, with the first air launches in 2005–2006. 48<br />

Flights <strong>of</strong> VESTA are also expected for trials <strong>of</strong> United Kingdom-led<br />

European ramjet technology development for the Meteor AAM. 49<br />

Germany began the development <strong>of</strong> ARMIGER missile in the<br />

mid-1990s to counter new antiaircraft defense systems currently in<br />

development. The missile is a Mach 3 IRR design using a boronbased<br />

solid-fueled DR (SFDR) sustainer and four midbody symmetrically<br />

positioned inlets. The missile has an asymmetric nose<br />

to accommodate the IRR seeker oriented at an angle relative to the<br />

flight direction to reduce friction heating that might confound the<br />

sensor. ARMIGER will use thrust controls, rather than control fins<br />

as a drag reduction measure. Whereas the missile is expected to enter<br />

service in 2008, technology advances are also expected to benefit<br />

the development <strong>of</strong> the Meteor missile.<br />

While France has investigated both SFRJ and LFRJ technology,<br />

Matra BAe Dynamics Aerospatiale (MBDA) seems to emphasize<br />

development <strong>of</strong> the later technology, believing the liquid-fueled designs<br />

have greater inherent performance potential. Initiated in 1996,<br />

France is today pursuing development <strong>of</strong> a low-cost LFRJ Rascal<br />

concept. The design uses high-pressure nitrogen to force fuel directly<br />

into the combustion chamber and computer-controlled on/<strong>of</strong>f<br />

injectors to control fuel flow. Component tests <strong>of</strong> this design were<br />

accomplished at Mach 2–2.5 at low altitude and Mach 3.2–3.7 at<br />

64,000 ft.<br />

Development <strong>of</strong> the Russian SS-N-26 Yakhont (Fig. 29), which<br />

began in 1998, was a significant step forward in terms <strong>of</strong> ramjet<br />

engine technology and threat to be countered. An engineering<br />

mockup was displayed at a recent air show <strong>of</strong> a land-attack derivative,<br />

Yakhont SS-NX-26 antiship cruise missile, with a reported<br />

extended range capability <strong>of</strong> 160 n mile for a total missile weight<br />

<strong>of</strong> 5,500 lbm. An IOC <strong>of</strong> 2003 is reported for the SS-N-26.<br />

An Indian/Russian joint development <strong>of</strong> the PJ-10 Brahmos was<br />

initiated in 1998. The Mach 2.8 missile is a modified derivative <strong>of</strong><br />

the Russian ramjet-powered SS-N-26 antiship missile. The propulsion<br />

system consists <strong>of</strong> an integral solid propellant booster and a<br />

liquid ramjet sustainer. Forward thrusters and aft jet vanes provide<br />

Fig. 29 Russian SS-N-26 operational missile (1998–today).<br />

Fig. 30 U.K. BVRAAM Meteor component tested missile (1999–<br />

today).<br />

the initial control to turn the missile to the direction <strong>of</strong> the target,<br />

whereas an inertial navigation system controls the missile during<br />

the midcourse and the terminal phase through an active radar seeker,<br />

with special electronic counter measures. Brahmos first flight tested<br />

in mid-2001 and is in progress today. A full-scale mockup <strong>of</strong> the<br />

Brahmos supersonic cruise missile was on display at the Indian DefExpo<br />

2002. A decision to begin commercial production <strong>of</strong> PJ-10<br />

Brahmos was announced September 2002. 59 The missile was recently<br />

showcased at India’s Republic Day celebrations in January<br />

2003, and the first naval launch was conducted in February 2003.<br />

Flight testing is ongoing, with the last test focusing on evaluating<br />

the guidance and fire control conducted October 2003.<br />

The BVRAAM/Meteor (Fig. 30) is a new concept in air-to-air<br />

weapons that employs solid propellant booster, advanced VFDR<br />

sustainer motor technology, and the latest electronics to deliver the<br />

required combat performance. Meteor will have the capability to<br />

engage multiple targets simultaneously, at greater range than current<br />

medium range AAM and in all weathers, day or night. It will<br />

complement Eur<strong>of</strong>ighter’s advanced short-range AAM capability,<br />

and it also is being developed to operate from Rafale and (assuming<br />

Swedish participation) Gripen aircraft. Meteor will be developed under<br />

a collaborative program involving the United Kingdom (lead),


Germany, Italy, and Spain (the Eur<strong>of</strong>ighter nations), France, and<br />

possibly Sweden. BVRAAM is designed to provide performance,<br />

particularly kinematic performance, several times that <strong>of</strong> existing<br />

MRAAMs.<br />

There is no active radar guided AAM in service with the Royal<br />

Air Force. Sky Flash is a semi-active missile and requires the launch<br />

aircraft to illuminate the target throughout the time <strong>of</strong> flight <strong>of</strong> the<br />

missile, which makes it vulnerable to counterattack. BVRAAM will<br />

give Eur<strong>of</strong>ighter the capability to engage multiple targets simultaneously,<br />

independent <strong>of</strong> parent aircraft maneuver, at greater range<br />

than AMRAAM and in all weathers, day or night. Following launch<br />

and in-flight target update, BVRAAM’s active radar seeker takes<br />

control and autonomously searches for and locks onto the target.<br />

Conventional rocket motor powered missiles rely on an initial<br />

boost phase to achieve the high speed required, followed by a coast<br />

phase to intercept the target. Latest generation, highly maneuverable<br />

aircraft are able to outrun conventional missiles at the extremes<br />

<strong>of</strong> their range. The VFDR airbreathing motor proposed for Meteor<br />

provides sustained power, following the initial boost, to chase and<br />

destroy the target. The missile’s computer and the seeker, which provides<br />

the missile’s ability to search, locate, and lock onto a target,<br />

will build on existing French technology, used in the Mica missile,<br />

to provide robust performance in the presence <strong>of</strong> electronic countermeasures.<br />

The BVRAAM/Meteor missile design, <strong>of</strong>fered by Matra BAE<br />

Dynamics and its consortium, was selected in May 2000 as most<br />

likely to meet U.K. needs over the life <strong>of</strong> the Eur<strong>of</strong>ighter aircraft,<br />

beating out the Raytheon Systems, Ltd., led consortium.<br />

PDE development activities in the United States have been pursued<br />

since the early 1990s. PDE missile system cost is predicted to<br />

be 30% <strong>of</strong> turbojet cost for a Mach 2.5 configuration with excellent<br />

fuel efficiency at high speed and to represent a potential 50% range<br />

increase at Mach 4. The PDE pro<strong>of</strong> <strong>of</strong> concept was demonstrated<br />

in single-pulse tests in the early 1990s. Flight-scale multiple-cycle<br />

tests with multiple combustors and a rotary valve were tested in the<br />

late 1990s. Ongoing efforts are proceeding into subsequent development<br />

phases.<br />

<strong>Ramjet</strong> Development: 2000–Today<br />

This era witnesses the continuation <strong>of</strong> international efforts to push<br />

the edge <strong>of</strong> the high-speed performance envelope. Table 4 also summarizes<br />

ramjet evolution in the era from 2000 to today and provides<br />

originating country, engine/vehicle names, development dates, performance,<br />

physical characteristics, and state <strong>of</strong> development. Countries<br />

engaged in ramjet development expanded significantly to include<br />

Japan, India, Sweden, Israel, and South Africa during this era.<br />

Scramjet development activities in this era were generally focused<br />

in the United States, Russia, Germany, Japan, and Australia.<br />

The U.S. Navy has supported SFRJ development activities for<br />

many years with the purpose <strong>of</strong> advancing the technology state <strong>of</strong><br />

the art. Such activities gained continued support in 2000 for continued<br />

development for a long-range, reduced-cost Mach 5–6 air- and<br />

surface-launched missile design.<br />

FRY 49<br />

Fig. 31 U.S. SSST flight-tested missile (2002–today).<br />

The People’s Republic <strong>of</strong> China has displayed a model at several<br />

air shows <strong>of</strong> a ramjet-powered air-launched missile similar in configuration<br />

to the French ASMP missile. The designation Ying-Ji 12<br />

suggests an antiship role similar to the French ANF missile.<br />

Following the cancellation <strong>of</strong> the SLAT in 1991, the U.S. Navy<br />

had to continue the quest for a replacement for the aging Vandal targets<br />

in the role <strong>of</strong> antiship cruise missile threat simulator. In the late<br />

1990s, the Navy evaluated the MA-31/AS-17 as an interim target.<br />

However, the MA-31 was not selected for large-scale production,<br />

and a new SSST missile (Fig. 31) was procured instead. Development<br />

began in 2000 on the GQM-163A Coyote nonrecoverable<br />

vehicle. The SSST is ground launched with a tandem Mk-70 solid<br />

propellant booster and uses an Aerojet (formerly Atlantic Research<br />

Corporation) MARC-R-282 VFDR ramjet sustainer that can reach<br />

speeds up to Mach 2.5 at sea level. Development testing was ongoing<br />

in 2003 with the vehicle successfully flighttested. Current plans<br />

call for production <strong>of</strong> six flight-test vehicles. If the tests are successful,<br />

up to 90 production targets may be ordered with an initial<br />

operational capability planned for mid-2004. 60<br />

The U.S. Navy initiated generic supersonic cruise missile<br />

(GSSCM) in 2002 as advanced cruise missile development for potential<br />

use in high-speed Tomahawk, high-speed strike or antiradiation<br />

applications. Low-drag LFIRR engine technology was proposed<br />

for the Mach 4–5 air- and surface-launched missile.<br />

The U.S. Navy is presently developing a propulsion system that<br />

will enhance the capabilities <strong>of</strong> an evolving antiradiation missile<br />

(ARM) system. The U.S. Navy initiated a four-year HSAD Project<br />

in late 2002 to address time-critical target requirements. The HSAD<br />

Project is directed toward increasing the range and average velocity<br />

<strong>of</strong> an advanced version <strong>of</strong> the HARM system, the higher speed ARM<br />

system. Based on the results <strong>of</strong> industry and government conceptual<br />

propulsion studies and a DOD ARM roadmap, the United States<br />

selected an IRR, VFDR propulsion system with tail-controlled steering.<br />

Three modified HARMs are to be built for captive carry and two<br />

air-launches from an F/A-18C/D in a technology demonstration. A<br />

formal three-year development program could start in 2006. 47,61<br />

Scramjet Development 1955–Today<br />

Scramjet Development: 1955–1990<br />

This era witnessed the beginning <strong>of</strong> scramjet development, with<br />

early combustion testing <strong>of</strong> ejector ramjets being initiated in the late<br />

1950s. Table 5 summarizes scramjet evolution in the era from 1955<br />

to 1990 and provides originating country, engine/vehicle names,<br />

development dates, performance, physical characteristics, and state<br />

<strong>of</strong> development. Significant developments in rocket boost and airbreathing<br />

propulsion systems that have occurred from midcentury<br />

onward greatly influenced the debate over hypersonic vehicle options<br />

and missions. The turbojet first flew in 1939, the ramjet in 1940,<br />

the high-performance large liquid-fueled rocket engine in 1943, and<br />

the practical man-rated reusable throttleable rocket engine in 1960.<br />

These dates serve as general milestones for numerous other developments,<br />

including supersonic afterburning turbojets, turboramjets,<br />

rocket- and turbine-based combined cycle engines, and scramjets.


50 FRY<br />

Fig. 32 First known freejet-tested scramjet in United States by Ferri<br />

(1963).<br />

Fig. 33 U.S. hypersonic research engine AIM (1966–1974).<br />

Other than the rocket, the ramjet has had the most direct effect<br />

on hypersonic design thinking. <strong>Ramjet</strong> technology matured rapidly<br />

following WWII. The emergence <strong>of</strong> scramjet propulsion cycle, successful<br />

ground-test demonstrations <strong>of</strong> liquid air collection under the<br />

U.S. Air Force Spaceplane program in the early 1960s, and the refinement<br />

<strong>of</strong> the airframe-integrated scramjet concept, all sparked<br />

great interest in scramjet propulsion for a wide range <strong>of</strong> hypersonic<br />

applications, interest which continues to the present day.<br />

The first known scramjet, fabricated and free-jet-tested, was a<br />

fixed-geometry engine model designed by A. Ferri (Fig. 32) in<br />

1960. Freejet tests <strong>of</strong> the complete engine were first performed in<br />

the GASL combustion-heated high-enthalpy blowdown tunnel in<br />

1963. These tests, integrated with aerodynamic tests, demonstrated<br />

that scramjets could be a viable propulsion system for hypersonic<br />

vehicles.<br />

The most extensive <strong>of</strong> the early (first generation) scramjet development<br />

programs in the United States was the NASA HRE<br />

program. 62 The HRE program, started in 1964, was crafted to<br />

develop and demonstrate scramjet technology. This program developed<br />

a flight-weight, variable-geometry, hydrogen-fueled and<br />

cooled scramjet engine designed to operate from Mach 4 to 7 and<br />

to be flight tested on the rocket-powered X-15A-2. The program<br />

progressed through component tests (inlet, combustor, and nozzle),<br />

flowpath tests, and flight-weight engine development and tests over<br />

aseven-year period. Following cancellation <strong>of</strong> the X-15 program<br />

in 1968, two axisymmetric engine models, the water-cooled structural<br />

assembly model (SAM) and the visually similar hydrogencooled<br />

aerothermodynamic integration model (AIM) (Fig. 33), were<br />

constructed and tested from 1971 to 1974. The two engines were<br />

tested to separately validate the flowpath operability and performance<br />

(AIM) and structural concepts (SAM). Flight-weight hardware<br />

was identified and developed to provide a complete scramjet<br />

engine system. During the HRE program, a total <strong>of</strong> 107 tests<br />

were performed on the two HRE dual-mode scramjet engines, es-<br />

tablishing an impressive and comprehensive database on inlet and<br />

combustor performance at Mach 5–7.<br />

France launched the ESOPE program in 1966, inspired at least in<br />

part by NASA’s HRE activity. The ESOPE effort was to demonstrate<br />

a dual-mode scramjet in a flight-test program at Mach 7. However,<br />

activities were ultimately limited to ground testing in two test series<br />

held between 1970 and 1972. Both the U.S. HRE and French ESOPE<br />

scramjet activities were terminated in favor <strong>of</strong> the development <strong>of</strong><br />

the IRR engine in support <strong>of</strong> missile applications.<br />

During the 1960s, as the propulsion needs <strong>of</strong> a new generation <strong>of</strong><br />

high-speed aircraft and reusable space transportation systems were<br />

addressed, integrating the rocket and the ramjet would now favor a<br />

specially designed integrated liquid-propellant rocket subsystem, an<br />

RBCC propulsion system. Improved performance benefits <strong>of</strong> rocket<br />

air-augmentation were attained by installing this rocket directly into<br />

the engine airflow path. Thrust and specific impulse for an ejector<br />

ramjet (ERJ) were approximately 15% higher than an equivalent<br />

rocket, rising to twice rocket levels at ramjet takeover conditions.<br />

The early Marquardt ERJ engine is a precursor <strong>of</strong> today’s RBCC<br />

family <strong>of</strong> engines. Although this class <strong>of</strong> combined cycle propulsion<br />

has yet to achieve flight-test status, extensive component testing<br />

has been conducted simulating a wide range <strong>of</strong> Mach number and<br />

altitude conditions. ERJ combustion testing in the 1960s includes<br />

16-in. gaseous hydrogen/air-fueled and 18-in. hydrogen/oxygen and<br />

JP4/hydrogen-peroxide-fueled configurations. Whereas hydrogen<br />

was preferred for space launch and hypersonic aircraft applications,<br />

conventional JP fuel with noncryogenic oxidizer was specified for<br />

conventional aircraft and missile applications. A fourth engine in<br />

this series, developed in 1968, used a fan-supercharged version <strong>of</strong><br />

the EJR configuration called the supercharged ERJ (SERJ). This engine<br />

was designed to power a Mach 4.5 high-performance aircraft<br />

and provide low fuel consumption at subsonic flight conditions.<br />

The supercharging fan was simulated in the inlet by direct-connect<br />

airflow control. By mid-1968, SERJ was tested over a range <strong>of</strong> simulated<br />

ramjet flight conditions up to Mach 3. Full-scale SERJ flight<br />

testing was proposed with a reengine <strong>of</strong> the rocket-powered X-15.<br />

The X-15 program was concluded in 1975, and these tests were<br />

never conducted.<br />

The U.S. Navy support <strong>of</strong> hypersonic propulsion began in the<br />

mid-1950s at JHU/APL in the form <strong>of</strong> the ERJ program. The intent<br />

<strong>of</strong> this effort was to demonstrate that both lift and thrust could be<br />

produced from the burning <strong>of</strong> fuels on the underside <strong>of</strong> wings when<br />

flying at supersonic or hypersonic speeds. The first-ever demonstration<br />

was subsequently conducted in 1958 <strong>of</strong> net positive thrust on a<br />

double wedge in a Mach 5 airstream. Following this early success, it<br />

was understood that much higher thrust and/or fuel specific impulse<br />

could be achieved. Two approaches were conceived by putting a<br />

cowl aft <strong>of</strong> the wedge or by ducting the flow through internal channels,<br />

much like other lower speed airbreathing cycles. However,<br />

unlike other cycles, the ducted scramjet also had to overcome many<br />

issues associated with hypersonic-speed flight, such as higher materials<br />

temperatures and heating rates, surface skin friction, and fuel<br />

ignition and kinetics. This work led to a follow on JHU/APL supersonic<br />

combustion ramjet missile (SCRAM) program.<br />

The U.S. Navy initiated SCRAM in 1961 to develop and demonstrate<br />

the technology necessary to prepare for the flight <strong>of</strong> an internally<br />

ducted scramjet-powered missile. Early studies showed that<br />

an internally ducted scramjet-powered missile could achieve powered<br />

ranges <strong>of</strong> several hundred miles when flying at Mach 8 at high<br />

altitude. The SCRAM and its components underwent considerable<br />

development work from the early 1960s to its termination in 1977.<br />

A large number <strong>of</strong> inlets, isolators, fuel injectors, liquid and gaseous<br />

fuels, ignition aids, and combustors were tested between Mach 3 and<br />

8. A 10-in. diam by 60-in. long three-module SCRAM freejet engine<br />

was tested in the 1968–1974 time frame from Mach 5.2 to 7.1 using<br />

liquid borane or mixtures <strong>of</strong> liquid hydrocarbon/borane fuels. Although<br />

the SCRAM program successfully demonstrated the technology<br />

necessary to proceed into flight testing, it had three unacceptable<br />

shortcomings: 1) logistically unsuitable pyrophoric and toxic liquid<br />

fuels and blends, 2) absence <strong>of</strong> sufficient room to house an active rf<br />

seeker, and 3) passive cooling requirements for the entire vehicle.


The U.S. JHU/APL devised the DCR as a successor to the<br />

SCRAM concept. Accordingly, the U.S. Navy initiated development<br />

<strong>of</strong> the Mach 4-6 hypersonic wide-area defense missile (Hy-<br />

WADM) or (WADM) in 1977 employing a DCR propulsion system<br />

with six forward fixed-geometry inlets, dual-feed subsonic gas generator<br />

combustion and four-feed subsonic/supersonic combustion.<br />

Engine component testing was conducted up to 1986, when funding<br />

for HyWADM was discontinued. DCR engine technology testing<br />

has continued intermittently since, with more activity resumed under<br />

the affordable rapid response missile demonstrator (ARRMD)<br />

and HyFly programs.<br />

In the mid-1980s, the U.S. NASP Program was formulated, with<br />

the objective <strong>of</strong> developing a SSTO “hypersonic combined-cycle airbreathing<br />

capable” 31 engine to propel a research vehicle, the X-30.<br />

The NASP program promise <strong>of</strong> flying a single-stage vehicle, powered<br />

by a combined cycle engine, which utilized scramjet operation<br />

to Mach 25 was aggressive, considering the state <strong>of</strong> the technology<br />

in 1986. Neither scramjet engines nor flowpaths had been tested<br />

above Mach 7. In addition, no credible, detailed analysis <strong>of</strong> scramjet<br />

performance, operability, loads, or structural approaches had ever<br />

been performed for flight past Mach 7. Also, what was good enough<br />

for Mach 7 vehicle operation was not refined enough for the lowthrust<br />

margin (energy from combustion vis-à-vis air kinetic energy)<br />

at double-digit flight Mach numbers. In other words, second generation<br />

scramjet technology was a good starting point, but considerable<br />

refinement and development were needed.<br />

Many significant contributions from the NASP program include<br />

hypervelocity scramjets, propulsion airframe integration methods/<br />

databases, capable structures, high-fidelity databases and design<br />

methods, and advancement <strong>of</strong> CFD methods. Extensive test programs<br />

were carried out for four basic scramjet engine concepts,<br />

providing comprehensive component and engine flowpath module<br />

databases for Mach numbers between 3 and 8. Large-scale flowpath<br />

testing up to Mach 12–14 was conducted but in pulse facilities with<br />

very short flow times. NASA LaRC performed over 1500 scramjet<br />

flowpath module tests to support the various contractor designs. Significant<br />

advancements were made in technology for hypervelocity<br />

scramjet design, design methods, test methodology (facilities and<br />

instrumentation), and database. However, scramjet technology for<br />

space access remained at least one generation behind scramjet technology<br />

for missiles, due primarily to the requirement for testing at<br />

such high Mach number. Additional discussion <strong>of</strong> NASP developments<br />

is beyond the scope <strong>of</strong> this paper.<br />

The Germans began development <strong>of</strong> Sänger II in the late 1980s as<br />

a proposed TSTO concept vehicle. It employed an airbreathing hypersonic<br />

first stage and delta wing second stage. The German Hypersonics<br />

Program and its Sänger II reference vehicle received most <strong>of</strong><br />

the domestic funding for spaceplane development in the late 1980s<br />

and early 1990s. Sänger II comprised a large hypersonic booster aircraft<br />

capable <strong>of</strong> Mach 4 cruise plus a small rocket-powered upper<br />

stage (HORUS) that could deliver people and cargo to low Earth<br />

orbit. The booster aircraft (to be powered by turboramjets) was<br />

designed for maximum commonality with a supersonic passenger<br />

transport (with a cruise range <strong>of</strong> 6000 n mile). Development would<br />

have been very costly and the program was canceled in 1994.<br />

Scramjet Development: 1990–2000<br />

Scramjet development came <strong>of</strong> age during this era, with the understanding<br />

<strong>of</strong> the technology that will soon enable scramjet-powered<br />

flight for the first time. Scramjet development in the 1990s initially<br />

focused on relatively near-term missile propulsion systems.<br />

Because <strong>of</strong> the basic difficulty <strong>of</strong> igniting and burning hydrocarbon<br />

fuels, missile designs employed methods to prepare the fuel or alternatively<br />

added highly energetic fuels or oxidizers for effective<br />

combustion. Several combustor designs have been investigated using<br />

various piloting techniques. The importance <strong>of</strong> active cooling<br />

for the hydrocarbon class <strong>of</strong> scramjet engines has been realized, as<br />

well as the value <strong>of</strong> using endothermic fuels. Scramjet development<br />

for space access continues to pickup momentum. Although axisymmetric<br />

engines were tested, the modular two-dimensional airframeintegrated<br />

supersonic combustion ramjet (SCRJ) engine emerged<br />

FRY 51<br />

as the candidate <strong>of</strong> choice. This configuration permits development<br />

in reasonable-size ground facilities and requires a relatively modest<br />

flight-test vehicle for a single module testing. We have come<br />

to accept the airframe-integrated lifting body as the standard vehicle<br />

configuration. In 1997, Billig 63 observed it prudent to examine<br />

the possibility <strong>of</strong> radical changes in engine flowpath and in turn<br />

overall vehicle configuration to produce a more effective vehicle.<br />

Some developments are exploring flowpath variations. Whereas the<br />

SCRJ is the key to airbreathing hypersonic flight, it is unlikely to<br />

provide efficient propulsion all <strong>of</strong> the way to orbital speeds. This<br />

era has witnessed the beginning development <strong>of</strong> combinations with<br />

mixed cycle engine designs. The hydrogen-fueled SCRJ will <strong>of</strong>fer<br />

acceptable performance to Mach 15. Hydrocarbon-fueled scramjet<br />

systems are being continued in exploration <strong>of</strong> propulsion technologies<br />

to speeds <strong>of</strong> Mach 8. Table 6 summarizes scramjet evolution in<br />

the era from 1990 to 2000 and provides originating country, engine/<br />

vehicle, development dates, performance, physical characteristics,<br />

and state <strong>of</strong> development. International scramjet development activities<br />

expanded dramatically during this era.<br />

Japan has pursued development <strong>of</strong> combined cycle engine technology<br />

since the late 1980s and early 1990s for flyback booster<br />

TSTO applications. ATREX is one element <strong>of</strong> the combined cycle<br />

airbreathing propulsion system being developed that is designed to<br />

give effective thrust to a spaceplane from sea level to altitude <strong>of</strong><br />

approximately 100,000 ft with a flight Mach number <strong>of</strong> 6 (Ref. 64).<br />

ATREX is a fan-boosted ramjet working on the expander cycle with<br />

three heat exchangers <strong>of</strong> hydrogen fuel, a precooler, an internal<br />

heat exchanger, and a regeneratively cooled combustor. The engine<br />

employs a tip turbine configuration that features compactness, light<br />

weight, and a variable-geometry inlet and plug nozzle, which allows<br />

operation under a wide range <strong>of</strong> flight conditions. Sea level static<br />

testing has been conducted since 1990 on ATREX-500, scaled down<br />

hardware with a fan inlet <strong>of</strong> 12-in. diameter and length <strong>of</strong> 87 in.<br />

Wind-tunnel testing has occurred since 1992 on an axisymmetric<br />

variable-geometry inlet, variable-geometry plug nozzles, and flying<br />

test bench hardware. Future flight testing <strong>of</strong> ATREX is planned.<br />

Supersonic combustion ramjet research activities in Russia have<br />

been pursued since the late 1950s. These activities began to accelerate<br />

in the 1980s and 1990s. In 1991, a decision was made to<br />

use SAMs to flight test the hypersonic ramjets for the first time.<br />

Kholod (Fig. 34) was developed as a first generation dedicated hypersonic<br />

flying laboratory, derived from the SA-5 (S-200 family)<br />

SAM, due to its trajectory being congenial to the hypersonic flighttest<br />

requirements. 65 An HRE-type E-57 hydrogen-fueled engine was<br />

used, consisting <strong>of</strong> an axisymmetric three-shock inlet, a coaxial regeneratively<br />

cooled combustion chamber, and low-expansion annular<br />

nozzle. The E-57 engine, which had a 9-in. diam cowl, is<br />

designed to fly at Mach 3.5–6.5 between altitudes <strong>of</strong> 50,000 and<br />

115,000 ft and remains attached to the booster rocket during the<br />

entire flight. 26,66,67 A total <strong>of</strong> seven testbed flights were performed,<br />

with four <strong>of</strong> these being cold-flow engine tests. These tests were not<br />

meant to demonstrate the viability <strong>of</strong> a specific engine applicable to a<br />

vehicle, but were intended to demonstrate in flight several technologies<br />

that were first developed in ground tests. These technologies<br />

included the following features: 1) dual-mode scramjet engine operation<br />

over a Mach number range <strong>of</strong> 3.5–6.5, including transition<br />

from subsonic combustion to supersonic combustion (mode transition);<br />

2) fuel-cooled engine structures; and 3) active control <strong>of</strong> fuel<br />

distribution and flow rate as a function <strong>of</strong> flight condition, as well<br />

as measured engine structural temperatures to allow demonstration<br />

<strong>of</strong> the first two technologies. Russia conducted the first flight test<br />

(Mach 5.35) in late 1991 and two joint Russian–France launches<br />

in late 1992 and early 1995. The second test reportedly achieved<br />

supersonic combustion conditions at Mach 5.6. In the third test, the<br />

engine failed to operate. 68 In late 1994, the NASA initiated a cooperative<br />

project to explore the scramjet operating envelope from the<br />

ram–scram, dual-mode operation below Mach 6 to the full supersonic<br />

combustion mode at Mach 6.5. To accomplish this objective,<br />

the higher heating loads required redesign <strong>of</strong> the combustor, active<br />

cooling system, and modifications; meanwhile, the increase to<br />

Mach 6.5 required modifications to the SA-5 booster performance.


52 FRY<br />

U.S. tests <strong>of</strong> the Russian pro<strong>of</strong>-test engine were planned, but never<br />

conducted for facility/model safety reasons. NASA engineers analyzed<br />

the final 1998 flight-test results. Although reasonable agreement<br />

was noted between ground- and flight-test data, 32 reportedly<br />

some uncertainty existed whether in-flight supersonic combustion<br />

conditions were achieved. Although these flight tests did not fully<br />

accomplish their original goals, they were a good first step, which<br />

helped build confidence that more ambitious flight tests could be<br />

accomplished, and just as important, they provided the first comparison<br />

between ground test and flight <strong>of</strong> dual-mode scramjet combustor<br />

data. Although still available, it is unlikely that further flight<br />

tests will be conducted with Kholod because more capable second<br />

generation hypersonic flying laboratories have become available.<br />

The Japanese National Aerospace Laboratory–KPL commenced<br />

design, fabrication, and testing <strong>of</strong> a side-wall compression-type<br />

scramjet engine in 1991 through 1994. 69 The flowpath design was<br />

based on the results <strong>of</strong> their research activities on the scramjet engine<br />

systems and components since the 1980s and had the objective<br />

<strong>of</strong> investigating component design and overall performance. A first<br />

generation hydrogen-fueled engine model E-1 was tested from 1994<br />

to 1999 at Mach 4–6 conditions. The majority <strong>of</strong> testing was conducted<br />

on heat-sink hardware, with limited water-cooled and liquid<br />

hydrogen-cooled testing. Two new facilities were built to support<br />

these activities; a freejet-type hypersonic propulsion wind tunnel<br />

(RJTF1) and a free-piston high-enthalpy shock tunnel (HIEST2)<br />

from 1994 to 1999.<br />

Design improvements were made, and a second generation E-2<br />

engine was fabricated with testing beginning in early 2000 and continuing<br />

today. Design changes focused primarily on the attainment<br />

<strong>of</strong> positive net thrust and improvement in combustion performance<br />

at Mach 8 conditions. The majority <strong>of</strong> the testing was conducted on<br />

water-cooled hardware, with limited liquid hydrogen-cooled and<br />

heat-sink testing. The first successful firing tests at Mach 12 conditions<br />

were conducted in 2002–2003. A plan to conduct scramjet<br />

flight testing is now underway to confirm engine performance under<br />

real flight conditions. A future milestone is to design, fabricate, and<br />

test the combined cycle engine based on the scramjet test results.<br />

French scramjet development activities reemerged in the late<br />

1980s with PREPHA, a jointly funded government–industry–<br />

university program, and subsequently reinitiated in 1992. PREPHA<br />

was aimed at developing a knowledge base on hydrogen-fueled dualmode<br />

ramjet technology for SSTO applications. 70 Generic waverider<br />

configurations were examined, which ultimately resulted in<br />

the design and test <strong>of</strong> CHAMIOS, a large-scale scramjet engine<br />

design that was ground tested at Mach 6.5. The hardware had a<br />

77 in. 2 (8 × 10 in.) entrance area and incorporated wall measurements<br />

and optical access. 71 Testing began in 1994 and continued<br />

through 1999, with further testing planned under other programs.<br />

Despite the potential <strong>of</strong> combined cycles for fully reusable space<br />

launchers, the PREPHA program ended in 1999. Follow-on developments<br />

were initiated to preserve the intellectual and material investment<br />

in the form <strong>of</strong> the WWR (1993–2003), JAPHAR (1997–2002)<br />

and Promethee (1999–2002) programs. 72<br />

Fig. 34 Russian Kholod first generation HFL (1991–1998).<br />

Fig. 35 Russian GELA/Raduga operational flight test vehicle (1995–<br />

today).<br />

The Russian Space Agency initiated comprehensive hypersonic<br />

research and development in the ORYOL (or OREL) program<br />

in 1993 to develop combined propulsion systems for advanced<br />

reusable space transportation, and includes SSTO and TSTO vehicle<br />

designs. 73 This program sought to focus over 40 years <strong>of</strong><br />

experience in Russian research and development in supersonic<br />

combustion. Activities are continuing with the development <strong>of</strong><br />

the Igla (shown subsequently) second generation hypersonic flying<br />

laboratory (HFL).<br />

GELA or Raduga (Fig. 35) was conceived as a Russian prototype<br />

for a new generation <strong>of</strong> hypersonic cruise missiles. The GELA<br />

testbed represents a second-phase effort on development work conducted<br />

by Russia between 1980 and 1985 (Ref. 74). The first phase<br />

dealt with Mach 3 ramjet propulsion systems developed from 1973<br />

to 1978, and ultimately used on the SS-N-22 and SA-6. A third phase<br />

<strong>of</strong> development seeks to build and test a Mach 6 missile, with a final<br />

phase seeking to achieve speeds up to Mach 8. The GELA missile<br />

was flight tested in 1994 from a Tu-22M (specially modified TU-95<br />

Bear) bomber at supersonic speed and boosted by a liquid propellant<br />

rocket motor and reportedly reached Mach 4.5. This vehicle is<br />

designed to conduct hypersonic scramjet research at speeds up to<br />

Mach 6.3 and 295,000 ft. It has been successfully launched upwards<br />

<strong>of</strong> 500 times. The configuration shown in Fig. 35 is an expendable<br />

vehicle.<br />

The U.S. Air Force has supported hypersonic engine technology<br />

activities since 1995 through the HyTech program. The fundamental<br />

objectives were to enable sustained hypersonic flight for missile<br />

or aircraft applications and to develop and demonstrate Mach 4–8<br />

hydrocarbon-fueled, actively cooled scramjet engine technology<br />

through component direct-connect and freejet testing. The Hypersonic<br />

<strong>Technology</strong> Development HyTech/HySET program evaluated<br />

concepts proposed by four engine contractors. Each contractor developed<br />

hydrocarbon-fueled scramjet concepts and databases, <strong>of</strong><br />

which two contractors were selected for the component demonstration<br />

in the first phase <strong>of</strong> the program. The Pratt and Whitney concept<br />

was selected to continue into the second phase under the HySET contract.<br />

This predominantly two-dimensional design heavily leveraged


20 years <strong>of</strong> hydrocarbon scramjet combustor development at UTRC.<br />

It also applied lessons learned from the hydrogen-fueled scramjet<br />

technology developed during NASP. The program is an applied research<br />

technology-based program to develop and demonstrate the<br />

structural durability and component performance <strong>of</strong> the inlet, isolator,<br />

combustor, and nozzle. Related technologies including inlets,<br />

composite leading edges, heat exchangers, and flameholding devices<br />

were all developed early in the program. These developments<br />

were followed by a full-scale performance demonstration <strong>of</strong> the engine<br />

design in a heavyweight uncooled copper engine. Integrated<br />

engine performance was demonstrated at Mach 4.5 and 6.5 conditions<br />

in freejet testing for this heavyweight copper engine, the<br />

performance test engine (PTE) under the HySET program. 75 The<br />

HyTech program has completed a major milestone in the successful<br />

testing <strong>of</strong> the world’s first flight-weight, fuel-cooled hydrocarbon<br />

ground demonstration engine number #1 (GDE-1). 76 This testing<br />

took the components demonstrated in PTE and integrated them into<br />

an engine and then demonstrated the total engine performance and<br />

durability with uncooled inlet leading edges and fuel-cooled panels<br />

that form the engine walls. Testing was similarly conducted at<br />

Mach 4.5 and 6.5 conditions. A follow-on engine, GDE-2 is now being<br />

constructed for testing in 2004. An added feature <strong>of</strong> the GDE-2<br />

is the variable-geometry inlet involving a pivoting inlet cowl flap.<br />

<strong>Technology</strong> from GDE-2 will flow into the NASA–U.S. Air Force<br />

X-43C and the DOD single engine demonstrator (SED).<br />

The United States has supported hypersonic flight research<br />

through NASA’s X-43A Hyper-X since 1995. The fundamental objectives<br />

have been to flight demonstrate the first integrated airframescramjet<br />

engine hypersonic vehicle and flight-validate key propulsion<br />

and related technologies. The program goal was to verify and<br />

demonstrate experimental, computational, and analytic design tools<br />

required for development <strong>of</strong> any hypersonic, airbreathing vehicle.<br />

Development focus was on an uncooled, hydrogen-fueled flowpath<br />

in support <strong>of</strong> hypersonic cruise aircraft and launch vehicles. Demonstration<br />

<strong>of</strong> critical technologies is planned in flight tests <strong>of</strong> the X-43A<br />

vehicle, two at Mach 7 and one at Mach 10. These vehicles are<br />

boosted to flight-test conditions using a modified Pegasus solid propellant<br />

booster, air-launched from a B-52 aircraft. Each vehicle is<br />

capable <strong>of</strong> about 10 s <strong>of</strong> powered flight. The program goal was<br />

to verify and demonstrate experimental techniques, computational<br />

methods, and analytical design tools required for the development<br />

<strong>of</strong> hypersonic, hydrogen-fueled, scramjet-powered aircraft. In addition,<br />

the program was to maximize the advancement <strong>of</strong> technology<br />

required for application <strong>of</strong> these engines to future aircraft and<br />

launch vehicles. The program included risk reduction and technology<br />

maturation tests <strong>of</strong> the Hyper-X scramjet in combustion, arc,<br />

and shock-heated facilities at Mach numbers <strong>of</strong> 7 and 10. Additional<br />

research tests conducted over the Mach 4–14 speed range characterized<br />

the Hyper-X engine concept over most <strong>of</strong> the scramjet operational<br />

range required for a launch vehicle. Over 800 tests (plus 700<br />

post-NASP, non-Hyper-X tests) were conducted and data compared<br />

with both analysis and facility-to-facility. The final test/verification<br />

<strong>of</strong> the Mach 7 X-43A propulsion-airframe-integrated scramjet was<br />

accomplished using a full-scale powered model plus flight engine in<br />

the LaRC 8-ft high temperature tunnel (HTT). 77 This flowpath testing<br />

provided many firsts. 28 Vehicle design was accomplished using<br />

tools developed in the NASP program.<br />

The first <strong>of</strong> the three X-43A flight-test vehicles was tested in June<br />

2001 (Fig. 36). Unfortunately, the booster vehicle and the X-43A<br />

had to be destroyed before separation and scramjet takeover. A<br />

mishap investigation board found the flight failure was unrelated<br />

to the propulsion system, but occurred because the vehicle’s control<br />

system was deficient in several analytical modeling areas, which<br />

overestimated the system’s margins. The next flight is planned for<br />

early 2004, which would be 54 years since the first ramjet-powered<br />

aircraft flight <strong>of</strong> Leduc.<br />

First conceived in 1993, the French government and industry are<br />

jointly funding a French–Russian program WRR, for reusable space<br />

launcher applications. The WRR prototype engine is a variablegeometry<br />

dual-mode ramjet that may involve either a rotating cowl<br />

(Promethée) or translating cowl (PIAF) inlet design. The vehicle is<br />

FRY 53<br />

a) NASA hyper-X flight test.<br />

b) B-52 with hyper-X release. c) Ignition <strong>of</strong> Pegasus rocket<br />

Fig. 36<br />

booster.<br />

U.S. first flight <strong>of</strong> X-43-A, June 2001 (1995–today).<br />

intended to fly at speeds from Mach 2 to 12, and uses kerosene fuel<br />

at the lower end <strong>of</strong> the speed range, then switches to hydrogen. Also<br />

<strong>of</strong> interest to this cooperative program is the study <strong>of</strong> integrating<br />

a detonation-based cycle and test methodology that allows examination<br />

<strong>of</strong> scale effects between small-scale flight testbed, 100-ft<br />

vehicle and full-scale space launcher. 71 The engine is a fuel-cooled<br />

design, which has involved testing <strong>of</strong> over 40 cooled-panel concepts<br />

to date. The cooperative program has faced some technological and<br />

budget problems, which produced delays, but strong interest exists<br />

for continuing the development. Subscale hardware (2 × 4 in. inlet<br />

entrance) testing at Mach 6 was conducted from 1995 to 1997. 78<br />

Testing <strong>of</strong> small-scale hardware (2 × 8 in. variable-geometry entrance)<br />

designated PIAF was initiated in late 2002, with full-scale<br />

hardware testing to follow. 49 Concept demonstration ground testing<br />

<strong>of</strong> prototype hardware (CHAMIOS-sized 8 × 10 in. inlet entrance)<br />

at Mach 3–6.5 is planned in the near future. Subscale (2-ton vehicle<br />

with 2 × 16 in. inlet entrance) flight testing is expected to follow<br />

around 2010 to demonstrate operation at Mach 1.5–4 and 8–12. Prototype<br />

(30-ton vehicle with 80 × 10 in. inlet entrance) flight testing<br />

is envisioned in the 2020 at Mach 1.5–12. An operational full-scale<br />

ramjet 500-ton vehicle with 2 × 33 ft entrance is foreseen in 2030.<br />

In the mid-1990s, Russia began openly discussing their development<br />

<strong>of</strong> AJAX, an innovative hypersonic vehicle concept. The<br />

concept fundamentally involves the capture <strong>of</strong> energy otherwise<br />

lost in flight at high Mach numbers and the recycling <strong>of</strong> this energy<br />

to increase the efficiency <strong>of</strong> the overall system. 79 The feasibility <strong>of</strong><br />

the approach depends on developing the systems required to capture<br />

energy from the flow and efficiently recycle it. The design consists<br />

<strong>of</strong> three elements: an MHD generator, a plasma airflow management<br />

system, and an endothermic fuel heat regeneration process.<br />

The following list gives some <strong>of</strong> the AJAX technologies.<br />

1) MHD generation <strong>of</strong> electrical power through deceleration <strong>of</strong><br />

the inlet flow, with the power generated used to provide the inlet<br />

stream ionization necessary to enable the MHD interactions to occur<br />

and the excess power at high Mach numbers available for other uses<br />

including producing a nonequilibrium plasma around the vehicle.<br />

2) Creation <strong>of</strong> a nonequilibrium cold plasma adjacent to the vehicle<br />

to reduce shock strength, drag, and heat transfer.<br />

3) Steam reforming <strong>of</strong> the hydrocarbon fuel through chemical<br />

regeneration, utilizing the endothermic nature <strong>of</strong> the steam reforming<br />

process to cool the vehicle and its engines, while producing<br />

methane and ultimately hydrogen onboard the vehicle for use in the<br />

high-speed scramjet engine cycle.<br />

The technologies associated with AJAX have been under investigation<br />

worldwide, and although there is much controversy regarding<br />

their effectiveness in modern hypersonic vehicles, it seems likely


54 FRY<br />

Fig. 37 France JAPHAR fixed dual-mode scramjet engine geometry<br />

(1997–2002).<br />

Fig. 38 Russian Igla second generation HFL (1999–today).<br />

that one or more <strong>of</strong> the technologies may be incorporated in vehicles<br />

designed for Mach > 6 flight.<br />

France began the development <strong>of</strong> JAPHAR in 1997 in cooperation<br />

with Germany as a follow-on to the PREPHA (French)<br />

and Sänger (German) programs to pursue hypersonic airbreathing<br />

propulsion research for reusable space launcher applications. 80<br />

Studies were organized around a waverider concept vehicle with a<br />

flight Mach between 4 and 8. JAPHAR employs a fixed-geometry<br />

dual-mode scramjet engine design (Fig. 37) developed based on previous<br />

hydrogen-fueled PREPHA work. Basic objectives included<br />

improvement and validation <strong>of</strong> numerical codes for aerothermodynamics<br />

and defining a flight-test vehicle. 72 Component hardware<br />

(4 × 4 in. entrance) testing was conducted from 2000 to 2002 simulating<br />

flight Mach numbers <strong>of</strong> 4.9, 5.8, and 7.6. French–German<br />

cooperation formally ended in 2001.<br />

In 1997, the DARPA initiated a program to develop ARRMD,<br />

as a low cost, air-launched Mach 4–8 missile. The program initially<br />

considered two hydrocarbon-fueled propulsion designs: the<br />

HyTech waverider two-dimensional engine and the axisymmetric<br />

DCR technology multiple-forward inlets. Both were potentially to<br />

make use <strong>of</strong> active cooling and a tandem solid propellant booster.<br />

The DCR engine was the design selected. Although the program<br />

was terminated in 2001 the DCR technology is being developed<br />

under the U.S. HyFly program, whereas the HyTech technology is<br />

being incorporated into the SED program.<br />

One <strong>of</strong> the Russian second generation flying laboratories was<br />

developed under the Igla (Fig. 38) program. 65 The first flight test<br />

was conducted in mid-1999, and a dummy Igla airborne testbed<br />

was displayed at the Russian Moscow airshow (MAKS) 1999 air<br />

show. The waverider vehicle is designed to achieve Mach 5–14 at<br />

an altitude <strong>of</strong> 82,000–164,000 ft. The Igla is boosted to supersonic<br />

speeds by the Rokot system; on separation, the flying testbed is<br />

injected into an operational trajectory for the scramjet engine, and<br />

the test vehicle is recovered by parachute. The testbed employs<br />

regeneratively cooled scramjet engine modules previously tested<br />

in ground-based facilities. Funding will dictate the future <strong>of</strong> this<br />

activity.<br />

The U.S. Navy has supported development <strong>of</strong> ATR technology<br />

under the Integrated High Performance Turbine Engine <strong>Technology</strong><br />

(IHPTET) program for potential application to air-launched Mach<br />

5 missiles since 1999. Engine development is focusing on the liquid<br />

air cycle engine turboramjet (ACETR) cycle with the purpose <strong>of</strong> re-<br />

ducing engine size and weight while maintaining theoretical performance.<br />

Although this engine cycle is immature and more complex<br />

when compared to alternative ramjet cycles, such an engine can<br />

provide increased loitering capability. Applications could include<br />

high-speed Tomahawk or other strike missile.<br />

The United States is investigating turbine-based propulsion systems<br />

for access to space under the revolutionary turbine accelerator<br />

(RTA)/TBCC project as part <strong>of</strong> the Next Generation Launch Technologies<br />

(NGLT) activities. 81 Present turbine propulsion systems<br />

can propel vehicles to Mach 3. These current systems are costly,<br />

require high maintenance, and have low durability. Near-term development<br />

goals <strong>of</strong> RTA will concentrate on turbine accelerators that<br />

will reach at least Mach 4 and provide dramatic increases in maintainability<br />

and operability through the use <strong>of</strong> advanced technologies.<br />

Studies suggest that the use <strong>of</strong> turbine propulsion can provide the potential<br />

for aircraftlike, space flight operations that may significantly<br />

reduce launch costs and improve safety. 82 During the initial phase <strong>of</strong><br />

RTA/TBCC, NASA GRC and general electric (GE) are designing a<br />

ground demonstrator engine for validation testing in fiscal year (FY)<br />

2006. The demonstrator is a turb<strong>of</strong>an ramjet, designed to transition<br />

from an augmented turb<strong>of</strong>an to a ramjet that produces the thrust required<br />

to accelerate the NGLT vehicle to Mach 4+. The initial flight<br />

test vehicle RTA-1 will demonstrate the basic TBCC concept <strong>of</strong> using<br />

a conventional turb<strong>of</strong>an to accelerate an access-to-space vehicle<br />

to Mach 3 and then transition to a ramjet mode designed to boost<br />

the vehicle to Mach 4+. Included in the testing is demonstration <strong>of</strong><br />

full-scale RTA enabling technologies along with reliability and durability<br />

<strong>of</strong> high-Mach turbine components, fuel, and control systems.<br />

In 2009, a second flight demonstrator RTA-2 will feature hardware<br />

at the scale <strong>of</strong> the vision propulsion system (VPS) product engine.<br />

RTA-2 combines the technology being developed in RTA-1 with<br />

advanced features from the DOD IHPTET and versatile affordable<br />

and advanced turbine engine (VAATE) programs and from NASA’s<br />

ultra efficient engine technology (UEET) program to meet the VPS<br />

goals <strong>of</strong> thrust to weight ratio, specific fuel consumption, specific<br />

impulse, safety, and cost. Together, initial and potential final configurations<br />

for RTA-1 and RTA-2 are expected to provide a technology<br />

readiness level-6 confidence in the key technology features needed<br />

to achieve the goals <strong>of</strong> the VPS. The program is also designed to<br />

meet the aggressive safety, cost, maintainability, and performance<br />

goals for the third generation RLV concept established by NASA. 81<br />

The French Promethée project, initiated in 1999, was aimed at<br />

developing fully variable-geometry endothermic hydrocarbon fuel<br />

dual-mode ramjet technology for military applications. The design<br />

is a generic air-to-surface missile able to fly at speeds <strong>of</strong> Mach 2–8<br />

and altitudes up to 130,000 ft. A full-scale combustion chamber<br />

was tested at simulated flight conditions <strong>of</strong> Mach 2–7.5. Under the<br />

original program, air-launched flight tests were planned at speeds <strong>of</strong><br />

Mach 4, 6, and 8 between 2009–2012. The flight-test vehicle was to<br />

be nonrecoverable. 49 The Promethée project was started by France<br />

to acquire first-hand knowledge <strong>of</strong> hydrocarbon-fueled dual-mode<br />

ramjet technology for military applications. 72 The program includes<br />

system studies and definition <strong>of</strong> a generic vehicle, design and optimization<br />

<strong>of</strong> a combustion chamber, and ground demonstration leading<br />

to flight validation. The French technology development has now<br />

progressed from direct-connect testing under the Promethée program<br />

to semifreejet and freejet testing in support <strong>of</strong> a new flight-test<br />

program called LEA initiated in 2003 and expected to run to 2012.<br />

India is conducting research and development activities for aerobic<br />

vehicle for advanced trans-atmospheric research (AVATAR),<br />

a reusable aerospace plane that is expected to be a 20-ton vehicle<br />

capable <strong>of</strong> 1000–2000 lbm payloads to low Earth orbit. A 3-ton<br />

subscale demonstrator vehicle AVATAR-M is under development.<br />

Scramjet Development: 2000-Today<br />

Scramjet technology developments are underway in many countries<br />

to capitalize on significant pay<strong>of</strong>fs that hypersonic speed and<br />

long range can provide. Table 6 also summarizes ramjet evolution<br />

in the era from 2000 to today and provides originating country,<br />

engine/vehicle names, development dates, performance, physical<br />

characteristics, and state <strong>of</strong> development.


Australia 83 conducted, with international support, the world’s first<br />

verified demonstration <strong>of</strong> supersonic combustion in a flight environment<br />

under the HyShot flight program. The demonstration results<br />

have received international endorsement for achieving supersonic<br />

combustion conditions. Two flight tests were conducted: one unsuccessful<br />

flight in late 2001 and a second successful flight in mid-<br />

2002, where supersonic combustion was observed. The model was<br />

a heat-sink copper scramjet configuration that retained the essential<br />

components <strong>of</strong> supersonic combustion, which consists <strong>of</strong> an intake<br />

and two combustion chambers. The thrust surfaces were removed<br />

for simplicity. Thus, strictly speaking, the model is not a scramjet<br />

and, hence, closely related to the gun-launched ram-accelerator<br />

work being pursued by many. The intake (4 in. width) is a simple<br />

wedge <strong>of</strong> 17-deg half angle followed by parallel combustion<br />

chambers (0.4 × 3 in.). Whereas the Kholod tests obtained dualmode<br />

scramjet combustor data over a range <strong>of</strong> Mach number, the<br />

HyShot experiment obtained not only supersonic combustion data at<br />

a single Mach, but a wide range <strong>of</strong> dynamic pressures, which were<br />

compared to ground-test data obtained over the same range. The<br />

model was boosted into a highly parabolic trajectory by a two-stage<br />

Terrier-Orion Mk70 rocket. The spent motor and model payload<br />

fell back to Earth, gathering speed such that between 120,000 and<br />

75,000 ft altitudes they were traveling at approximately Mach 7.6.<br />

The program demonstrated that 1) an understanding <strong>of</strong> supersonic<br />

combustion gained from shock-tunnel ground tests is sufficient to<br />

design a simple supersonic combustor that will operate in flight and<br />

2) the test approach is a cost effective means to undertake hypersonic<br />

flight testing.<br />

Among the second generation flying testbeds currently explored<br />

by Russia is the use <strong>of</strong> a MIG-31 interceptor as a launch aircraft<br />

for exploring conditions from Mach 2 to 10 (Ref. 65). Unlike the<br />

axisymmetric Kholod, the study <strong>of</strong> two-dimensional hypersonic engines<br />

are more easily integrated into the body <strong>of</strong> an aircraft. Use<br />

<strong>of</strong> this aircraft seeks to overcome the shortcomings common to existing<br />

ground-based experimental facilities, which cannot provide a<br />

full simulation <strong>of</strong> all <strong>of</strong> the conditions <strong>of</strong> complex engine exposure<br />

to aerodynamic and heat loadings at speeds <strong>of</strong> Mach 6–8. This variant<br />

<strong>of</strong> flying laboratory allows research at conditions <strong>of</strong> Mach 2–10,<br />

altitudes <strong>of</strong> 50,000–130,000 ft, dynamic pressures <strong>of</strong> 1.4–30 lbf/in. 2 ,<br />

and an operational time for the scramjet <strong>of</strong> 40 s. The recoverable<br />

scramjet test module is air launched using a modified SA-10 solid<br />

propellant booster up to Mach 10 conditions in a ballistic trajectory.<br />

Test history revealed at the 1999 Moscow Air Show indicated<br />

100 min <strong>of</strong> direct-connect testing and 60 min <strong>of</strong> freejet testing.<br />

The U.S. Navy and DARPA initiated the HyFly (Fig. 39) program<br />

in early 2002 to mature and demonstrate, in-flight, DCR<br />

scramjet propulsion technology to enable hypersonic long-range<br />

missiles. 84 This four-year demonstrator project evolved from existing<br />

Navy hypersonic efforts and from DARPA’s ARRMD program.<br />

Direct-connect combustor testing begun in 2000 is continuing<br />

for Mach 3–6.5 and will continue for some time. Freejet tests at<br />

Mach 6–6.5 conditions were successfully completed in mid-2002<br />

<strong>of</strong> the JP-10-fueled, uncooled DCR engine fully integrated into a<br />

flight-representative aeroshroud or missile body. Future development<br />

will include additional direct-connect and wind-tunnel ground<br />

tests <strong>of</strong> the full-scale flight-weight engine into 2004. Subscale ballistic<br />

flights <strong>of</strong> the engine mounted on a sounding rocket are in progress<br />

with Mach 4 and Mach 6 cruise flight tests anticipated in 2004 and<br />

2005–2006, respectively. The Navy hypersonic propulsion developments<br />

are generally focused exclusively on missile applications,<br />

whereas the U.S. Air Force/NASA developments are aimed at both<br />

reusable platforms as well as one-time use vehicles.<br />

NASA initiated the integrated system test <strong>of</strong> an airbreathing<br />

rocket (ISTAR) program in 2002 with the objective to flight test a<br />

self-powered vehicle to more than Mach 6 by the end <strong>of</strong> the decade,<br />

which would demonstrate all modes <strong>of</strong> RBCC engine operation.<br />

Fig. 39 U.S. HyFly flight demonstration vehicle (2002–today).<br />

FRY 55<br />

The engine employs a hydrocarbon-fueled liquid rocket system for<br />

initial acceleration, with a ramjet that ignites at about Mach 2.5,<br />

followed by conversion to scramjet operation around Mach 5. The<br />

engine will have the capability <strong>of</strong> accelerating to Mach 7. The strutjet<br />

RBCC engine design was selected for ground demonstration and<br />

subsequent flight testing. Ground testing <strong>of</strong> a flight-weight, fuelcooled<br />

engine flowpath is scheduled to begin in 2006. The scramjet<br />

engine designs examined are expected to provide performance 15%<br />

above conventional rockets during the initial boost phases <strong>of</strong> the<br />

flight. Funding for the ISTAR flight test program was cancelled in<br />

late 2003, but ground testing is continuing.<br />

In 2002 the U.S. initiated X-43B as a follow-on development<br />

activity to X-43A Hyper-X. The program involves a flight demonstration<br />

<strong>of</strong> reusable RBCC or TBCC advanced vehicles in a notional<br />

10-min flight in the Mach 0.7–7 range. The RBCC engine<br />

was expected to be a strutjet, hydrocarbon-fueled and cooled design.<br />

The TBCC engine is expected to trace it roots to the HyTech<br />

and RTA technology bases. In late 2003 NASA decided to focus<br />

upon the TBCC flowpath for the X-43B and has renamed the effort<br />

the reusable combined cycle flight demonstrator (RCCFD). Flight<br />

testing is planned for the 2010 time frame.<br />

In 2003, France initiated LEA, a new flight-test demonstration <strong>of</strong><br />

a high-speed dual-mode ramjet propelled vehicle (Fig. 40) at flight<br />

conditions <strong>of</strong> Mach 4–8. The program is planned to demonstrate the<br />

feasibility <strong>of</strong> a positive aeropropulsive balance. 85 It will allow definition<br />

and flight validation <strong>of</strong> ground development methodologies for<br />

predicting aeropropulsive balance and required design margins. The<br />

final propulsion system used may be a fixed geometry (JAPHAR)<br />

or variable geometry (Promethée or PIAF). It is not planned to alter<br />

the geometry in-flight if a variable geometry design is selected.<br />

The expected flight consists <strong>of</strong> an aircraft release, acceleration <strong>of</strong><br />

the flight vehicle by a solid booster to the desired Mach number,<br />

booster separation, vehicle stabilization, and autonomous flight for<br />

20–30 s. Six flights are planned between 2010 and 2012. Semifreejet<br />

and freejet inlet testing activities are currently ongoing in support<br />

<strong>of</strong> the planned flight-test program. 86,87<br />

The technology developed by the United States in the HyTech<br />

program will next flow into two newly funded flight-test programs,<br />

the NASA–U.S. Air Force X-43C flight demonstration begun in<br />

1999 and the DOD SED begun in 2003. The SED (Fig. 41) program<br />

Fig. 40 France LEA concept flight test vehicle (2003–today).<br />

Fig. 41 U.S. endothermically fueled SED flight demonstrator (2003–<br />

2006).


56 FRY<br />

will demonstrate the operation <strong>of</strong> the endothermically fueled scramjet<br />

engine in flight using a single flowpath and fixed-geometry,<br />

self-starting inlet. It is based on the use <strong>of</strong> the HyTech waverider<br />

two-dimensional engine configuration and features an army tactical<br />

missile system (ATACMS) solid propellant booster to accelerate the<br />

waverider test vehicle to a scramjet takeover Mach number <strong>of</strong> 4.5<br />

with the scramjet engine further accelerating the vehicle to Mach<br />

7. Five test flights are planned starting in 2005–2006 to precede the<br />

X-43C flight tests.<br />

The most ambitious application <strong>of</strong> the U.S. HyTech technology is<br />

the joint NASA–U.S. Air Force X-43C program. This program seeks<br />

to develop waverider flight-test vehicles that will accelerate from<br />

Mach 5.5 to 7 using three flowpaths developed from the HyTech<br />

hydrocarbon-fueled scramjet engine. Each flight-weight fuel-cooled<br />

flowpath will feature the variable-geometry cowl flap <strong>of</strong> the GDE-<br />

2 engine. Like the X-43A, the X-43C will be air launched from<br />

a L-1011 and accelerated by a Pegasus solid propellant booster<br />

to Mach 7+. Ground testing will include single- and multi-module<br />

nose-to-tail propulsion airframe integration demonstrators and flight<br />

clearance engines. Three flight tests are scheduled, beginning in<br />

2007.<br />

General State <strong>of</strong> <strong>Ramjet</strong>, Scramjet, and Mixed<br />

Cycle <strong>Technology</strong><br />

Flight-Demonstrated <strong>Technology</strong><br />

The state <strong>of</strong> the flight-demonstrated ramjet technology base in<br />

1980 was summarized by Thomas87 and is shown here in Fig. 42.<br />

Fewer operational ramjet systems were known to have existed in<br />

the open literature. Those actually in existence included subsonic<br />

U.S. Navy drones, Bomarc, Talos, and D-21, the Bloodhound and<br />

Sea Dart in the United Kingdom, the SE 4400, VEGA, and CT-41<br />

in France, the C-101 in the People’s Republic <strong>of</strong> China, and the<br />

SA-4, SA-6, and SS-N-19 in Russia. Development activities were<br />

principally focused in the United States and Russia, with the French,<br />

Germans and Chinese beginning ramjet missile activities. Thomas<br />

noted that the major milestones <strong>of</strong> the period from 1960 to 1980<br />

were the development <strong>of</strong> low-altitude short-range missile and the<br />

conception and demonstration <strong>of</strong> the IRR for missile applications.<br />

The evolution <strong>of</strong> ramjet and scramjet technology from 1918 to<br />

today was reviewed in the preceding sections and summarized in<br />

Tables 2–6. An appreciation for the advances made may be obtained<br />

by reviewing Tables 2–6 and noting that the vast majority <strong>of</strong> the test<br />

database is seen to exist between Mach 3 and 8.<br />

An interesting comparison may be made by superimposing the<br />

1980 technology database shown in Fig. 43 and ramjet and scramjet<br />

system performance data provided in Tables 2–6 on the typical<br />

airbreathing flight envelope discussed earlier (Fig. 2). A direct comparison<br />

<strong>of</strong> the 1980 and 2003 ramjet technology bases is provided<br />

in Fig. 43. Worldwide development activities have advanced the<br />

demonstrated upper speed range considerably: Operational systems<br />

matured from Mach 3.5 to 4.5, prototype engines matured from<br />

Mach 4.5 to 7, advanced technology flight tests have matured from<br />

Mach 5.5 to 10, and ground-test feasibility testing has matured from<br />

Mach 8 to 10–18. The operational range <strong>of</strong> dynamic pressure has also<br />

Fig. 42 State <strong>of</strong> flight-demonstrated ramjet technology base in 1980.<br />

Fig. 43 Comparison <strong>of</strong> 1980 and 2004 flight-demonstrated ramjet<br />

technology bases.<br />

correspondingly broadened. The state <strong>of</strong> the 2004 technology base,<br />

shown in Fig. 43, shows a dramatic expansion over the last 20 years.<br />

Airbreathing engines and their supporting rocket technology for<br />

space access face the following potential issues, many <strong>of</strong> which<br />

also apply to military applications 89 : thrust limitations, especially<br />

around Mach = 1, demonstration <strong>of</strong> efficient engine operation over<br />

broad flight Mach number and altitude range, structural and propellant<br />

fractions for both airbreathing and rockets, logistics <strong>of</strong> fuel<br />

(H2 vs kerosene), materials thermal environment, and engine design<br />

complexity.<br />

Hypersonic airbreathing engines in combination with other engine<br />

cycles are the most promising for affordable access to space and<br />

high-speed cruise. High-temperature materials and efficient propulsion<br />

performance over a broad Mach number and altitude range<br />

are keys to successful development <strong>of</strong> these vehicles. A further examination<br />

<strong>of</strong> today’s state <strong>of</strong> engine cycle maturity concludes this<br />

discussion.<br />

<strong>Ramjet</strong>/Scramjet Engine <strong>Technology</strong><br />

As a result <strong>of</strong> research over the last 50 years, considerable advances<br />

have been made in airbreathing propulsion technology. Particularly<br />

significant is the state <strong>of</strong> flight testing at Mach > 7 that<br />

can be used to validate ground testing and the rapid advancements<br />

in liquid hydrocarbon scramjet engine technology. Although the<br />

X-43A is on the verge <strong>of</strong> demonstrating scramjet performance at<br />

Mach 7, currently only rocket propulsion has demonstrated flight<br />

performance at high Mach numbers beyond Mach > 6.<br />

The maturity <strong>of</strong> ramjet/scramjet airbreathing propulsion technology<br />

resides at the actual system flight-test level at Mach 3, decreasing<br />

monotonically to system prototype at Mach 5, decreasing further<br />

to component test in a simulated flight environment at Mach 8–10<br />

and continuing at this level out to approximately Mach 15. Newer<br />

airbreathing mixed engine cycles, such as RBCC and TBCC, are<br />

at the component test maturity at Mach < 7. Airframe development<br />

maturity follows slightly behind propulsion. Critical issues continue<br />

to be focused on airframe thermal performance, propulsion performance,<br />

overall flight efficiency, and development <strong>of</strong> flight-weight<br />

subsystems. It is now instructive to turn our attention to current<br />

research needs.<br />

Summary<br />

Advances in ramjet technology over the past century have been remarkable,<br />

involving dramatic advances in flight-demonstrated technologies.<br />

The road to discovery has not been without its distractions,<br />

which include world events as well as inconsistent support for the<br />

burgeoning technology by resident governments. Initial motivation<br />

began with the desire for propelling advanced aircraft, followed<br />

by missile technology, and now encouraged by the development <strong>of</strong><br />

reusable Earth-to-orbit vehicles that employ airbreathing engines for<br />

at least a portion <strong>of</strong>, if not the entire, flight envelope. Since the turn<br />

<strong>of</strong> the century, the expansion <strong>of</strong> the operational envelope for ramjets<br />

has been dramatic and range from the beginning notions <strong>of</strong> flight<br />

to testing <strong>of</strong> engine designs that approach previously inconceivable<br />

speeds.


In 2004, we are approaching speed ranges entirely unheard <strong>of</strong><br />

to the Wright brothers in 1903 or to Leduc, who in 1957 spoke <strong>of</strong><br />

600-mph limits for aircraft applications. Currently, sufficient knowledge<br />

exists to challenge the upper speed limits <strong>of</strong> pure ramjets<br />

(M ∼ 6–8), dual-mode scramjets (M ∼ 14), and pure or rocketassisted<br />

scramjets (M ∼ 20). Nevertheless, the greatest knowledge<br />

base exists in the Mach 3–7 flight range. Strides have been taken in<br />

recent years to expand this understanding to include the upper region<br />

to orbital speeds, as well as the lower region to static conditions.<br />

It cannot escape notice that since 1990 the international activity in<br />

ramjet and scramjet missile development has increased noticeably.<br />

Very strong international ramjet and scramjet capabilities are being<br />

created through many significant ongoing developments. Whereas<br />

propulsion approaches to future requirements continue to evolve,<br />

it appears that the combination <strong>of</strong> current programs has served to<br />

revive and reinvigorate a new generation <strong>of</strong> industrial and military<br />

capability.<br />

<strong>Ramjet</strong> technology has matured to a high state <strong>of</strong> readiness for<br />

military applications. Greater stand<strong>of</strong>f ranges and reduced time to<br />

target are consistently mentioned in conjunction with future missile<br />

requirements. <strong>Ramjet</strong> or scramjet solutions certainly provide<br />

the kinematic properties desired, but remaining factors such as affordability,<br />

payload integration, inlet packaging, and development<br />

risk all play important roles in a selection process. The attractive<br />

performance attributes <strong>of</strong> ramjet-powered missiles have been available<br />

for over 50 years; however, limited applications have come to<br />

being, at least within the United States Recent advances in integral<br />

booster design may help reduce many system-level concerns, and<br />

advances in targeting and information technology may create the<br />

need for the added range that ramjet propulsion can supply. Hybrid<br />

or mixed cycle ramjet technology is developing to support future<br />

supersonic and high-speed transport. Scramjet technology has matured<br />

considerably in the last 15 years and promises to open this<br />

new century <strong>of</strong> flight with the first flight <strong>of</strong> a scramjet-propelled vehicle<br />

with true potential for enabling space access. Still, substantial<br />

advances are required to support military and reusable launch vehicle<br />

applications. Advanced developments, such as PDE and MHD<br />

technologies, are progressing and show great promise for expanding<br />

the potential <strong>of</strong> high-speed airbreathing vehicles.<br />

On a fundamental level, our understanding is maturing on turbulence<br />

and its effects at higher speeds, on wall shear and heat transfer,<br />

boundary-layer separation and reattachment, fuel injection and mixing,<br />

and chemical kinetics and combustion dynamics in an engine.<br />

CFD is becoming an increasingly important tool in understanding<br />

these fundamental processes, combined with an expanding database<br />

for validating physical and chemical models used. Strides have also<br />

been taken to expand the engineering design database on mixed<br />

cycle engine performance at low and high speeds to complement<br />

the extensive existing database in the Mach 3–7 range. Certainly,<br />

opportunities in research and development continue to exist and<br />

will do so well into this second century <strong>of</strong> ramjet history. To an<br />

aerospace engineer entering our field today, the outlook is bright<br />

and the future exciting. The authors are reminded <strong>of</strong> similar excitement<br />

surrounding the activity leading up to and culminating with<br />

the initial successful landing <strong>of</strong> NASA’s Apollo 13 on the moon in<br />

1969. The authors challenge the international community to maintain<br />

focus and resolve for a consistent effort to realize the promises<br />

for airbreathing flight into this new century as mankind continues<br />

to push back the frontiers <strong>of</strong> flight and seek a better understanding<br />

for our place in the universe.<br />

It is apparent that airbreathing technology has matured to occupy<br />

an important place in the propulsion field. The authors have<br />

only touched the surface <strong>of</strong> the technology. Its future importance,<br />

although hopeful, cannot be foreseen based on past history. It is a<br />

truism that technology feeds on itself, that work in one area <strong>of</strong>ten<br />

is quickly applicable in an entirely different area. We can all contribute<br />

more effectively to using this process and explaining it to<br />

our communities and funding sources to justify its existence. We<br />

must be cautioned against the casual hindsight judgment <strong>of</strong> the idea<br />

whose time had come. More than once, participants were convinced<br />

wrongly that airbreathing’s time had come, were tempted to assume<br />

FRY 57<br />

that a favorite project or system’s existence was inevitable, and found<br />

that this belief contributed to the failure <strong>of</strong> our own purposes. <strong>Ramjet</strong><br />

and scramjet propulsion’s time shall come only after solutions to<br />

challenging technical problems are resolved and the need is clear.<br />

Acknowledgments<br />

The author is deeply appreciative to R. Waesche for our many<br />

years <strong>of</strong> fruitful collaboration and our valuable discussions, based<br />

on his many years <strong>of</strong> experience in the development <strong>of</strong> ramjet technology,<br />

which contributed to this review. The author wishes to acknowledge<br />

the significant contributions made to the state <strong>of</strong> the<br />

art <strong>of</strong> ramjet and scramjet technology by individuals worldwide.<br />

The author gratefully acknowledges constructive reviews by Stuart<br />

Blashill, U.S. Navy; Parker Buckley, Robert Mercier, and F. D. Stull<br />

(retired), U.S. Air Force Research Laboratory; and Chuck McClinton,<br />

NASA Langley Research Center. The author welcomes receipt<br />

<strong>of</strong> additional historical material or references, which can be incorporated<br />

in a future review.<br />

References<br />

1Sloop, J., “Liquid Hydrogen as a <strong>Propulsion</strong> Fuel, 1945–1959,” Forward<br />

by Fletcher, J. C., NASA SP-4404, NASA Scientific and Technical<br />

Informative Office, Washington, DC, April 1977.<br />

2Avery, W. H., “Twenty-Five Years <strong>of</strong> <strong>Ramjet</strong> Development,” Jet<strong>Propulsion</strong>,<br />

Vol. 25, No. 11, 1955, pp. 604–614.<br />

3Thomas, A. N., “Some Fundamental Aspects <strong>of</strong> <strong>Ramjet</strong> <strong>Propulsion</strong>,” Jet<br />

<strong>Propulsion</strong>, Vol. 27, No. 4, 1957, pp. 381–385.<br />

4Courtesy <strong>of</strong> the U.S. Air Force <strong>Propulsion</strong> Directorate, circa 1960s.<br />

5Waltrup, P. J., “Hypersonic Airbreathing <strong>Propulsion</strong>: <strong>Evolution</strong> and Opportunities,”<br />

CP-428, AGARD, April 1987.<br />

6Zucrow, M. J., Aircraft and Missile <strong>Propulsion</strong>, John Wiley and Sons,<br />

NY, Vols. 1 and 2, 1958.<br />

7Dugger, G. L. (ed.), <strong>Ramjet</strong>s, AIAA selected reprints, Vol. 6, AIAA, New<br />

York, 1969.<br />

8 “The Pocket <strong>Ramjet</strong> Reader,” United Technologies Corp., Chemical Systems<br />

Div., 1978.<br />

9Seebass, A., Murthy, S., and Curran, E. (eds.), High Speed Flight <strong>Propulsion</strong><br />

Systems, Progress in Astronautics and Aeronautics, Vol. 137, AIAA,<br />

Washington, DC, 1991.<br />

10Billig, F. S., and Dugger, G. L., “The Interaction <strong>of</strong> Shock Waves and<br />

Heat Addition in the Design <strong>of</strong> Supersonic Combustors,” Proceedings <strong>of</strong><br />

Twelfth (International) Symposium on Combustion, Combustion Inst., Pittsburgh,<br />

PA, 1969, pp. 1125–1134.<br />

11Billig, F. S., Dugger, G. L., and Waltrup, P. J., “Inlet–Combustor Interface<br />

Problems in Scramjet Engines,” Proceedings <strong>of</strong> the 1st International<br />

Symposium on Airbreathing Engines, June 1972.<br />

12Waltrup, P. J., and Billig, F. S., “Prediction <strong>of</strong> Precombustion Wall Pressure<br />

Distribution in Scramjet Engines,” Journal <strong>of</strong> Spacecraft and Rockets,<br />

Vol. 10, No. 9, 1973, pp. 620–622.<br />

13Waltrup, P. J., and Billig, F. S., “Structure <strong>of</strong> Shock Waves in Cylindrical<br />

Ducts,” AIAA Journal, Vol. 11, No. 9, 1973, pp. 1404–1408.<br />

14Heiser, W. H., and Pratt, D. T., Hypersonic Airbreathing <strong>Propulsion</strong>,<br />

AIAA Education Series, AIAA, Washington, DC, 1994.<br />

15Edwards, T., “Liquid Fuels and Propellants for Aerospace <strong>Propulsion</strong>:<br />

1903–2003,” Journal <strong>of</strong> Power and <strong>Propulsion</strong>, Vol. 19, No. 6, 2003.<br />

16Wittenberg, H., “Some Fundamentals on the Performance <strong>of</strong> <strong>Ramjet</strong>s<br />

with Subsonic and Supersonic Combustion,” TNO Prins Maurtis Lab., 2000.<br />

17Ferri, A., “Review <strong>of</strong> Problems in Application <strong>of</strong> Supersonic<br />

Combustion—Seventh Lanchester Memorial Lecture,” Journal <strong>of</strong> the Royal<br />

Aeronautical Society, Vol. 68, No. 645, 1964, pp. 575–597.<br />

18Courtesy <strong>of</strong> the U.S. Air Force <strong>Propulsion</strong> Directorate, circa 1990s.<br />

19Escher, W. J. D., and Flornes, B. J., “A Study <strong>of</strong> Composite <strong>Propulsion</strong><br />

Systems for Advanced Launch Vehicle Applications,” Marquardt Corp.,<br />

Rept. 25194, NASA Contract NAS 7-377, Vols. 1–7, Sept. 1966.<br />

20Curran, E. T., “Scramjet Engines—The First Forty Years,” 13th International<br />

Symposium on Air Breathing Engines, ISABE Paper 97-7005, Sept.<br />

1997.<br />

21Curran, E. T., “Emerging <strong>Propulsion</strong> Technologies for the Next <strong>Century</strong>,”<br />

AIAA Wright Brothers Lecture, World Aviation Congress and Exposition,<br />

Oct. 1999.<br />

22Waltrup, P. J., White, M. E., Zarlingo, F., and Gravlin, E. S., “History <strong>of</strong><br />

U.S. Navy <strong>Ramjet</strong>, Scramjet, and Mixed-Cycle <strong>Propulsion</strong> Development,”<br />

AIAA Paper 96-3152, July 1996.<br />

23 “Back to the Future,” video presentation, Kaiser Marquardt Corp., 1997.<br />

24Kuentzmann, P., and Falempin, F., “<strong>Ramjet</strong>, Scramjet and PDE—An<br />

Introduction,” Encyclopedia <strong>of</strong> Physical Science and <strong>Technology</strong>, 3rd ed.,<br />

Academic Press, Vol. 13, 2002.


58 FRY<br />

25VanWie, D. M., “Bumblebee and Beyond: <strong>Propulsion</strong> Advancements<br />

at the Johns Hopkins University Applied Physics Laboratory,” International<br />

Symposium on Air Breathing Engines, ISABE Paper 2003-1010,<br />

Sept. 2003.<br />

26Andrews, E. H., “Scramjet Development and Testing in the United<br />

States,” AIAA/NAL 10th International Space Planes and Hypersonic Systems<br />

and Technologies Conf., AIAA Paper 2001-1927, April 2001.<br />

27Escher, W. J. D., “A U.S. History <strong>of</strong> Airbreathing/Rocket Combined-<br />

Cycle (RBCC) <strong>Propulsion</strong> for Powering Future Aerospace Transports, with<br />

A Look Ahead to the Year 2020,” International Symposium on Air Breathing<br />

Engines, ISABE Paper 1999-7028, 1999.<br />

28McClinton, C. R., Andrews, E. H., and Hunt, J. L., “Engine Development<br />

for Space Access: Past, Present and Future,” International Symposium<br />

on Air Breathing Engines, ISABE Paper 2001-1074, Jan. 2001.<br />

29Ferri, A., “Review <strong>of</strong> Scramjet <strong>Technology</strong>,” Journal <strong>of</strong> Aircraft,Vol. 5,<br />

No. 1, 1968, pp. 3–10.<br />

30Anderson, G., McClinton, C., and Weidner, J., “Scramjet Performance,”<br />

Scramjet <strong>Propulsion</strong>, Progress in Astronautics and Aeronautics, Vol. 189,<br />

AIAA Reston, VA, 2000, Chap. 6.<br />

31Escher, W. J. D., “The Synerjet Engine,” Synerjet for Earth/Orbit<br />

<strong>Propulsion</strong>: Revisiting the 1966 NASA/Marquardt Composite (Airbreathing/<br />

Rocket) <strong>Propulsion</strong> Study, SAE Progress in <strong>Technology</strong> Series, PT-54, Society<br />

<strong>of</strong> Automotive Engineers, Warrendale, PA, 1966, Chap. 3.<br />

32McClinton, C. R., Rausch, V. L., Sitz, J., and Reukauf, P., “Hyper-X<br />

Program Status,” 10th International Space Planes and Hypersonic Systems<br />

and Technologies Conf., AIAA, April 2001.<br />

33Hueter, U., and McClinton, C., “NASA’s Advanced Space Transportation<br />

Hypersonic Program,” 11th AIAA/AAF International Space Planes and<br />

Hypersonic Systems and Technologies Conf., AIAA 2002-5175, Sept. 2002.<br />

34Sänger, E., and Bredt, I., “A Ram-Jet Engine for Fighters,” NACA<br />

TM 1106, Dec. 1947; translation <strong>of</strong> “Uber ein Loreneintreib fur Strahjager,”<br />

Deutsche Luftfahrtforschung, Untersuchungen und Mitteilungen,<br />

Rept. 3509.<br />

35Escher, W. J. D., “Back to the Future, Our <strong>Ramjet</strong>/Rocket <strong>Propulsion</strong><br />

Heritage–A Kaiser Marquardt Historical Review,” 1997.<br />

36Escher, W. J. D., and Foreman, K. M., “Major <strong>Ramjet</strong> Programs in the<br />

U.S.,” International Symposium on Air Breathing Engines, ISABE Paper<br />

2003-10-72, Sept. 2002.<br />

37Courtesy <strong>of</strong> the U.S. Air Force <strong>Propulsion</strong> Directorate, circa 1970s.<br />

38Janes’s Defense Weekly, 10Sept. 2001.<br />

39Marguet, R., Berton, P., and Hirsinger, F., “L’etude du Statoreacteur<br />

Supersonique et Hypersonique en France de 1950 e 1974 (Application aux<br />

Moteurs Combines Aerobies), CP-479, AGARD, Dec. 1990.<br />

40Directory <strong>of</strong> U.S. Military Rockets and Missiles, 2003.<br />

41Webster, F., and Bucy, J., “ASALM PTV Chin Inlet <strong>Technology</strong><br />

Overview,” AIAA Paper 79-1240, June 1979.<br />

42Webster, F., “Evaluation <strong>of</strong> the ASALM-PTV <strong>Propulsion</strong> System Flight<br />

Data Correlation Results and Analysis Techniques,” AIAA Paper 1981-1606,<br />

July 1981.<br />

43Washington Times, 25Sept. 2001.<br />

44Petit, B., “<strong>Ramjet</strong>s and Scramjets,” Encyclopedia <strong>of</strong> Physical<br />

Science and <strong>Technology</strong>, 3rd ed., Academic Press, Vol. 13, 2002,<br />

pp. 867–884.<br />

45Wilson, R., Limage, C., and Hewitt, P., “The <strong>Evolution</strong> <strong>of</strong> <strong>Ramjet</strong> Missile<br />

<strong>Propulsion</strong> in the U.S. and Where we are Headed,” AIAA Paper 96-3148,<br />

July 1996.<br />

46Dunfee, D. D., and Hewitt, P. W., “Worldwide Developments in <strong>Ramjet</strong>-<br />

Powered Missiles (A Situational Review),” CPIA Publ. 630, Vol. 1, Chemical<br />

<strong>Propulsion</strong> Information Agency, Columbia, MD, 1995.<br />

47Hewitt, P. W., “Worldwide Developments in <strong>Ramjet</strong>-Powered<br />

Missiles—An Update from 1995,” CPIA Publ. 713, Vol. 1, Chemical <strong>Propulsion</strong><br />

Information Agency, Columbia, MD, April 2002.<br />

48 “World Missiles Briefing,” Teal Group Corp., Feb. 2003.<br />

49Jane’s Missiles and Rockets, 1,Aug. 2002.<br />

50 “World Missiles Briefing,” Teal Group Corporation, March 2003.<br />

51Janes Missiles and Rockets, Vol. 7, No. 10, Oct. 2002, p. 8.<br />

52U.S. Naval Inst. Military Database, July 2002.<br />

53Barrie, D., “Precision Pursuit,” Aviation Week and Space <strong>Technology</strong>,8<br />

Sept. 2003.<br />

54Flight International Magazine, 10–16 Dec. 1997.<br />

55International Defense Review, 1Feb. 1996, p. 53.<br />

56Jane’s Missiles and Rockets, Vol. 3, No. 10, Oct. 1999.<br />

57Wall, R., and Fulghum, D., “Vympel Battles to Maintain Air-to-Air<br />

Missile Lead,” Aviation Week and Space <strong>Technology</strong>, 8Oct. 2001.<br />

58U.S. Naval Inst., Military Database, June 2001.<br />

59 “BrahMos Production to Start in 2004,” Jane’s Missiles and Rockets,<br />

Vol. 6, No. 11, 2002.<br />

60Directory <strong>of</strong> U.S. Military Rockets and Missiles, 2003.<br />

61 Wall, R., “JSF, Better Performance Drives HARM Upgrade,” Aviation<br />

Week and Space <strong>Technology</strong>, 23Dec. 2002, p. 28.<br />

62 Andrews, E. H., and Mackley, E. H., “NASA’s Hypersonic Research<br />

Engine Project–A Review,” AIAA 1993-2323, June 1993.<br />

63 Billig, F. S., “Overview <strong>of</strong> <strong>Propulsion</strong> Performance,” AGARD Conf. on<br />

Future Aerospace <strong>Technology</strong> in the Service <strong>of</strong> the Alliance, April 1997.<br />

64 Sato, N., Balepin, V., Harda, K., Kobayashi, H., Hatta, H., Kogo, Y.,<br />

Kashiwagi, T., Mizutani, T., Omi, J., Hamabe, K., Tomike, J., Minami, R.,<br />

and Kretschmer, J., “Development Study on ATREX Engine,” AIAA Paper<br />

96-4553, Nov. 1996.<br />

65 Romashkin, A., “Russian Hypersonic Technologies,” Cosmonautics,<br />

2002, pp. 60–62.<br />

66 Hicks, J., “International Efforts Scram into Flight,” Aerospace America,<br />

June 1998, pp. 28–33.<br />

67 McClinton, C., Roudakov, A., Semenov, V., and Kopchenov, V., “Comparative<br />

Flowpath Analysis and Design Assessment <strong>of</strong> an Axisymmetric<br />

Hydrogen Fueled Scramjet Flight Test Engine at a Mach Numbert <strong>of</strong> 6.5,”<br />

AIAA Paper 96-4571, Nov. 1996.<br />

68 Roudakov, A., Semenov, V., Kopchenov, V., and Hicks, J., “Future Flight<br />

Test Plans <strong>of</strong> an Axisymmetric Hydrogen-Fueled Scramjet Engine on the<br />

Hypersonic Flying Laboratory,” AIAA Paper 96-4572, Nov. 1996.<br />

69 Chinzei, N., “Scramjet Engine Tests at NAL-KPL, Japan,” International<br />

Symposium on Air Breathing Engines, ISABE Paper 2003-1171, Sept. 2003.<br />

70 Bonnefond, T., Falempin, F., and Viala, P., “Study <strong>of</strong> a Generic<br />

SSTO Vehicle Using Airbreathing <strong>Propulsion</strong>,” AIAA Paper 96-4490,<br />

Nov. 1996.<br />

71 Chevalier, A., Levine, V., Bouchez, M., and Davidenko, D., “French–<br />

Russian Partnership on Hypersonic Wide Range <strong>Ramjet</strong>s,” AIAA Paper 96-<br />

4554, Nov. 1996.<br />

72 Serre, L., and Falempin, F., “PROMETHEE: The French Military Hypersonic<br />

<strong>Propulsion</strong> Program Status in 2002,” AIAA Paper 2002-5426, Oct.<br />

2002.<br />

73 Lanshin, A., and Sosounov, V., “Russian Aerospace Combined <strong>Propulsion</strong><br />

Systems Research and Development Program (Oryol-2-1): Progress<br />

Review,” AIAA Paper 96-4494, Nov. 1996.<br />

74 Janes’s Review, “Russia Still Finds Cash for ‘Silver Bullet’ Solutions,”<br />

Vol. 2, No. 10, 1 Oct. 1995.<br />

75 Norris, R. B., “Freejet Test <strong>of</strong> the AFRL HySET Scramjet Engine Model<br />

at Mach 6.5 and 4.5,” 37th AIAA Joint <strong>Propulsion</strong> Conf., July 2001.<br />

76 Boudreau, A. H., “Status <strong>of</strong> the U.S. Air Force HyTech Program,” Inernational<br />

Symposium on Air Breathing Engines, ISABE Paper 2003-1170,<br />

Sept. 2003.<br />

77 McClinton, C. R., Rausch, V. L., Sitz, J., and Reukauf, P., “Hyper-X<br />

Program Status,” 10th International Space Planes and Hypersonic Systems<br />

and Technologies Conf., AIAA, April 2001.<br />

78 Bouchez, M., Falempin, F., Levine, V., Avrashkov, V., and<br />

Davidenko, D.,“French–Russian Partnership on Hypersonic Wide Range<br />

<strong>Ramjet</strong>s: Status in 2002,” AIAA Paper 2002-5254, Oct. 2002.<br />

79 Gurijanov, E. P., and Harsha, P. T., “AJAX: New Directions in Hypersonic<br />

<strong>Technology</strong>,” AIAA Paper 96-5211, Nov. 1996.<br />

80 Dessornes, O., and Scherrer, D., “Weighing <strong>of</strong> the JAPHAR Dual Mode<br />

<strong>Ramjet</strong> Engine,” AIAA Paper 2002-5187, Oct. 2002.<br />

81 Shafer, D., and McNelis, N., “Development <strong>of</strong> a Ground Based Mach<br />

4+ Revolutionary Turbine Accelerator <strong>Technology</strong> Demonstrator (RTATD)<br />

for Access to Space,” International Symposium on Air Breathing Engines,<br />

ISABE Paper 2003-1125, Sept. 2003.<br />

82 Bradley, M., Bowcutt, K., McComb, J., Bartolotta, P., and McNelis,<br />

N., “Revolutionary Turbine Accelerator (RTA) Two-Stage-to-Orbit (TSTO)<br />

Vehicle Study,” AIAA Paper 2002-3902, July 2002.<br />

83 Paull, Allan, Alesi, H., and Anderson, S., “HyShot Flight Program and<br />

How it was Developed,” AIAA Paper 2002-4939, Oct. 2002.<br />

84 Kandebo, S. W., “New Powereplant Key to Missile Demonstrator,” Aviation<br />

Week and Space <strong>Technology</strong>, 2Sept. 2002, pp. 56–59.<br />

85 Falempin, F., and Serre, L., “LEA Flight Test Program,” AIAA Paper<br />

2002-5249, Oct. 2002.<br />

86 Serre, L., and Minard, J.-P., “French Potential for Semi-Free Jet Test <strong>of</strong><br />

an Experimental Dual Mode <strong>Ramjet</strong> in Mach 4–8 Conditions,” AIAA Paper<br />

2002-5448, Oct. 2002.<br />

87 Falempin, F., Goldfeld, M., Nestoulia, R., and Starov, A., “Experimental<br />

Investigation <strong>of</strong> Variable Geometry Inlet at Mach Number from 2 to 8,”<br />

International Symposium on Air Breathing Engines, ISABE Paper 2003-<br />

1162, Sept. 2003.<br />

88 Thomas, A. N., “New-Generation <strong>Ramjet</strong>s—A Promising Future,” Astronautics<br />

and Astronautics, June 1980, pp. 36–41, 71.<br />

89 Waltrup, P. J., “The Application <strong>of</strong> Hypersonic Air Breathing <strong>Propulsion</strong><br />

to Space Access and High Speed Vehicles in the U.S.,” Johns Hopkins<br />

Univ./Applied Physics Lab. Eugen Sänger Lecture for Year 2000/2001,<br />

Columbia, MD, Oct. 2001.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!